Molecular mechanisms of cell–cell spread of intracellular bacterial

advertisement

Review

Cite this article: Ireton K. 2013 Molecular mechanisms of cell – cell spread of intracellular bacterial pathogens. Open Biol 3: 130079.

http://dx.doi.org/10.1098/rsob.130079

Received: 9 May 2013

Accepted: 23 June 2013

Subject Area: microbiology/cellular biology

Keywords:

Listeria

,

Shigella

,

Rickettsia

, actin-based motility, protrusions, cell junctions

Author for correspondence:

Keith Ireton e-mail: keith.ireton@otago.ac.nz

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016

rsob.royalsocietypublishing.org

Molecular mechanisms of cell – cell spread of intracellular bacterial pathogens

Keith Ireton

Department of Microbiology and Immunology, University of Otago, Dunedin, New Zealand

1. Summary

Several bacterial pathogens, including Listeria monocytogenes , Shigella flexneri and

Rickettsia spp., have evolved mechanisms to actively spread within human tissues.

Spreading is initiated by the pathogen-induced recruitment of host filamentous

(F)-actin. F-actin forms a tail behind the microbe, propelling it through the cytoplasm. The motile pathogen then encounters the host plasma membrane, forming a bacterium-containing protrusion that is engulfed by an adjacent cell. Over the past two decades, much progress has been made in elucidating mechanisms of

F-actin tail formation.

Listeria and Shigella produce tails of branched actin filaments by subverting the host Arp2/3 complex. By contrast, Rickettsia forms tails with linear actin filaments through a bacterial mimic of eukaryotic formins.

Compared with F-actin tail formation, mechanisms controlling bacterial protrusions are less well understood. However, recent findings have highlighted the importance of pathogen manipulation of host cell–cell junctions in spread.

Listeria produces a soluble protein that enhances bacterial protrusions by perturbing tight junctions.

Shigella protrusions are engulfed through a clathrin-mediated pathway at ‘tricellular junctions’—specialized membrane regions at the intersection of three epithelial cells. This review summarizes key past findings in pathogen spread, and focuses on recent developments in actin-based motility and the formation and internalization of bacterial protrusions.

2. Introduction

Rapid microbial dissemination (‘spread’) within key host organs is a critical step in many infectious diseases. In the case of some intracellular bacterial pathogens, spread between human cells involves a phenomenon called actin-based

motility (ABM) [1,2]. Bacteria that exhibit ABM include the enteric pathogens

Listeria monocytogenes and Shigella flexneri , and select species of the arthropodborne genus Rickettsia

[1,2]. The hallmark of ABM is subversion of the host

actin cytoskeleton to stimulate bacterial motility within a human cell. This intracellular motility ultimately leads to microbial spread between host cells.

Cell–cell spread and other crucial steps in the intracellular life cycles of

Listeria , Shigella and Rickettsia

are depicted in figure 1 [1–7]. After internalization

into human cells, bacteria are initially enclosed in host membranous structures called phagosomes (step 1). Within 30–60 min, phagosomes are destroyed through bacterial factors, allowing microbes access to the cytosol (step 2). Cytoplasmic bacteria replicate (step 3) and become decorated with host-derived actin filaments (step 4). Recruitment of F-actin is due to bacterial surface proteins that stimulate polymerization of actin monomers. The actin filaments organize into tail-like structures that produce ABM by propelling bacteria through the cytoplasm. Motile bacteria form protrusions derived from the host plasma membrane (step 5). These protrusions are ultimately internalized by surrounding host cells, resulting in bacteria encased in double membranous vacuoles (step 6).

&

2013 The Authors. Published by the Royal Society under the terms of the Creative Commons Attribution

License http://creativecommons.org/licenses/by/3.0/, which permits unrestricted use, provided the original author and source are credited.

Figure 1.

Steps in the intracellular life cycles of the bacterial pathogens

Listeria monocytogenes

,

Shigella flexneri

and

Rickettsia

spp. (1) internalization of bacteria into host cells, (2) destruction of phagosomes and access of bacteria to the host cytosol, (3) replication in the cytosol, (4) ABM, (5) formation of bacterial protrusions, (6) engulfment of protrusions and (7) dissolution of the double membranous vacuole. The process of cell – cell spread comprises steps 4 – 7.

The initially infected cell that generates bacterial protrusions is coloured in yellow, whereas an adjacent cell internalizing a protrusions is blue. The plasma membranes of these cells are coloured differently in order to illustrate the origin of the two membranes in the vacuole resulting from protrusion engulfment (step 6). ‘AJC’ denotes the apical junctional complex—a structure composed of tight junctions and adherens junctions.

Bacterial enzymes destroy these vacuoles, liberating microbes and allowing infection of new human cells (step 7). For the purpose of this review, we define cell–cell spread as steps 4–7, starting with ABM and ending with escape from the double membranous vacuole. Throughout the spreading process, bacteria are encased in the plasma membrane of host cells. For this reason, spreading is thought to allow intracellular colonization of host tissues while shielding the pathogen from immune

responses involving antibody or complement [6,8].

Over the past approximately 25 years, much progress has been made in understanding the molecular mechanisms of several aspects of the intracellular life cycles of Listeria and

Shigella . In particular, bacterial and/or host proteins that mediate internalization, phagosomal escape, cytoplasmic replication, actin polymerization and destruction of the

double membranous vacuole have been identified [3 –7,

9–33]. In many cases, how bacterial and/or human factors

act at a molecular level to achieve these steps is partly understood. Research with Rickettsia has progressed more slowly compared with work with Listeria or Shigella , mainly because of difficulties in bacterial genetic analysis. However, recent results have shed light on modes of Rickettsia internalization

and ABM [2,4,34–42]. Compared with steps 1–4 of the

Listeria , Shigella or Rickettsia life cycles, the formation and engulfment of protrusions (stages 5 and 6) have proved more difficult to elucidate. Before 2005, it was unclear whether protrusions are generated simply as a passive con-

sequence of ABM [43] or if instead mechanisms exist that

act after bacterial-directed actin polymerization to directly

govern protrusion formation [44]. In addition, whether

uptake of protrusions involves active participation of the human cell was not understood. Recent findings identifying bacterial and host proteins that control protrusion formation

[44 –46] and human factors needed for engulfment of protrusions [47,48] represent important first steps in understanding

post-ABM stages of bacterial spreading.

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016

This review will cover molecular aspects of cell–cell spread of Listeria , Shigella and Rickettsia , focusing on ABM (step 4), protrusion formation (step 5) and protrusion internalization (step 5).

Special emphasis will be given on findings obtained over the past 8 years. Other steps in the intracellular life cycles of these

pathogens have been recently reviewed [2,4,6,7,34,49].

2

3. Actin-based motility

ABM of Listeria , Shigella or Rickettsia is promoted by bacterial surface proteins that are structurally distinct and stimulate

actin polymerization through different means [1,2,5,49–51].

The Listeria protein ActA and Shigella protein IcsA both activate a mammalian actin polymerization machinery known as the

Arp2/3 complex. The precise mechanisms of Arp2/3 activation by ActA or IcsA differ. The requirement for Arp2/3 in ABM of

Listeria or Shigella indicates that these bacteria subvert an existing actin polymerization pathway in the human cell. By contrast, the Rickettsia protein Sca2 appears to directly stimulate assembly of actin filaments independently of Arp2/3 or other host factors. Sca2 may act as a functional mimic of a class of eukaryotic proteins called formins. In §3.1, I provide a brief summary of the mechanisms of actin assembly promoted by

Arp2/3 or formins. For more extensive discussions of actin polymerization, the reader is referred to several recent reviews

[50,52–54]. In §3.2, I describe actin assembly induced by

Listeria , Shigella or Rickettsia , with emphasis on recent findings.

3.1. Mechanisms of actin polymerization in eukaryotic cells

3.1.1. Actin filament assembly and function

Actin is present in a monomeric form, or as filaments derived

from the polymerization of several actin monomers [50,54].

Actin filaments have a defined polarity, which determines the overall direction of filament growth. Actin monomers complexed with ATP tend to add on to the barbed (plus) end of an existing filament. ATP hydrolysis occurs in the filament, resulting in actin-ADP that ultimately dissociates from the pointed (minus) end of the filament. Addition of actin monomers on to the barbed end of a filament is enhanced by the protein profilin, which stimulates the exchange of ATP for

ADP on monomers [55,56]. In cells, actin filaments are often

dynamic, assembling or disassembling in response to external stimuli such as growth factors or extracellular matrix com-

ponents [50,54]. The regulated assembly or disassembly of

F-actin plays critical roles in many cellular processes. For example, actin polymerization helps remodel membranes, probably by generating force at actin–membrane interfaces

[50,57]. This membrane remodelling function of F-actin con-

tributes to many essential processes, including cell motility, cytokinesis, endocytosis and vesicular trafficking from the

endoplasmic reticulum or Golgi apparatus [52,58]. Depending

on the process, force generation may involve simply actin polymerization or the combined action of F-actin and myosin

to produce contractility [58]. The controlled disassembly of

actin filaments also impacts many important biological

events, including regulated exocytosis [58] and the engulfment of particles through phagocytosis [59].

The first step in the de novo assembly of an actin filament

is the formation of actin dimers or trimers [50,54]. This

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016 process, termed ‘nucleation’, is rate-limiting in vitro . In cells, several proteins exist that accelerate nucleation, thereby stimulating actin polymerization. These ‘nucleators’ fall into three general classes: the Arp2/3 complex, formin proteins

and WH2 domain-containing nucleators [50,53,60– 65].

Arp2/3- and formin-mediated nucleation are relevant to known mechanisms of bacterial ABM and are therefore discussed later. WH2 domain nucleators will not be covered in this review.

3.1.2. Nucleation of actin filaments by the Arp2/3 complex or formins

Arp2/3 is an evolutionarily conserved complex of seven pro-

teins [13,50,52,66– 71]. Two of the seven components (Arp2

and Arp3) have structural similarity to monomeric actin

[72]. The Arpc1 component has a WD40 domain that forms

a seven-bladed beta propeller. The remaining components

(Arpc2, Arpc3, Arpc4 and Arpc5) do not exhibit significant structural similarity to other known proteins. The Arp2/3 complex stimulates polymerization of a new actin filament from the side of an existing (‘mother’) filament, resulting in

a Y-shaped branched actin structure [60,73,74] (figure 2

a (i)).

Studies involving electron tomography suggest that Arpc2 and Arpc4 contact the mother actin filament, whereas Arp2 and Arp3 interact with pointed end of the nascent filament.

[75]. The Arp2 and Arp3 components are thought to form a

dimer on the side of the mother filament, serving as the

first subunits of the new actin filament [75,76]. Thus, the

Arp2/3 complex may stimulate actin polymerization by mimicking an actin dimer, whose formation is normally the rate-limiting step in filament assembly.

By itself, the Arp2/3 complex is inefficient in promoting

F-actin assembly. Efficient actin polymerization requires activation of the Arp2/3 complex by ‘nucleation-promoting

factors’ (NPFs) [60,77–79]. One of the best characterized

NPFs is neuronal Wiscott–Aldrich syndrome protein, or

‘N-WASP’ [80]. N-WASP uses a region called a WCA domain

to activate Arp2/3 (figure 2 a

(i)). This domain interacts with

the Arp2/3 complex [78], inducing conformational changes

that bring the Arp2 and Arp3 components into close proximity

and render the complex competent for nucleation [81–84]. In

addition to activating the Arp2/3 complex, N-WASP also binds and delivers actin monomers to the nucleation machinery. The WCA domain interacts with monomeric actin, and a

proline-rich region binds actin complexed with profilin [85,86].

N-WASP is itself subject to complex regulation (figure

2

a

(ii)) [50,80]. In the absence of cellular stimuli, N-WASP is

autoinhibited due to intramolecular interactions that mask

the activity of the WCA domain [78,87]. The protein WIP

(WASP-interacting protein) stabilizes the inactive confor-

mation of N-WASP [88]. In response to growth factors or

other stimuli, autoinhibition of N-WASP is relieved through interactions with several cellular factors including the activated (GTP-bound) form of the small GTPase Cdc42, the lipid phosphatidylinositol 4,5-bis phosphate and Src Homology 3 domains from the signalling proteins Toca-1, Nck

or Grb2 [78,89–94]. In addition to these regulatory inter-

actions, activation of N-WASP is also promoted by serine phosphorylation of its WCA domain, which increases the

affinity of this domain for the Arp2/3 complex [95].

In contrast to the branched actin structures produced by the Arp2/3 complex, formin proteins nucleate the assembly

of linear actin filaments (figure 2 b

) [50,52,61,62]. The core

elements common to all formins are the two ‘formin homology’ domains FH1 and FH2. Formins dimerize to form a ring-like structure that associates with the plus end of an

actin filament [96– 98]. The FH2 domains of formin dimers

stimulate filament nucleation, probably by stabilizing actin

dimers [99]. This domain also promotes elongation of actin

filaments by preventing factors known as ‘capping proteins’

from halting polymerization [62,100,101]. The formin FH1

domain binds actin –profilin complexes, an action thought to increase the local availability of actin monomers for

addition to the filament plus end [102,103].

At least 15 mammalian formins exist [50,53]. These proteins

fall into seven different classes based on the FH2 domain amino acid sequence and the presence of additional domains that

regulate FH1 and FH2 function [50,104]. The best-understood

formin class is the diaphanous-related formins, comprising the

proteins mDia1, mDia2 and mDia3 [50,53]. The FH1 and FH2

domains are located in the carboxyl-terminal region of Dia proteins. This region also contains a ‘diaphanous autoinhibitory domain’ (DAD) that regulates the actin polymerization activity of the FH2 domain. The DAD controls this activity by interacting with a ‘diaphanous inhibitory domain’ (DID) and ‘GTPasebinding domain’ (GBD) located in an amino-terminal region of

Dia proteins. Binding of the DAD to the DID and GBD results in

autoinhibition [105,106]. Upon cellular stimulation via engage-

ment of cell surface receptors, autoinhibition is relieved by association of activated forms of the small GTPases Cdc42 or

RhoA with the GBD of Dia proteins [107–109].

The Arp2/3 complex controls a variety of essential processes in mammalian cells, including endocytosis and membrane trafficking between the endoplasmic reticulum and Golgi

apparatus [50,52,57]. Both Arp2/3 and diaphanous-related for-

mins promote the formation of membrane extensions, called lamellipodia or filopodia, that drive cell motility, formation of

cell–cell junctions and phagocytosis [50,52]. Given the ability

of the Arp2/3 complex and Dia proteins to produce actin filaments that generate force and remodel cellular membranes, it is not surprising that many intracellular microbial pathogens have evolved mechanisms to exploit these two pathways of

actin polymerization [2]. In §3.2, I explain how the bacteria

Listeria and Shigella manipulate Arp2/3 or N-WASP in order to promote ABM. I also describe recent results indicating that Rickettsia stimulates intracellular motility by producing a bacterial mimic of eukaryotic formins. In §4, I provide further examples of bacterial subversion of Arp2/3 or formin function during protrusion formation leading to cell–cell spread.

3

3.2. Bacterial stimulation of actin assembly

3.2.1.

Listeria

Listeria monocytogenes is a Gram-positive food-borne pathogen capable of causing serious infections culminating in abortions

or meningitis [3,110]. Cell–cell spread of

Listeria is thought to be critical for disease, based on observations that bacterial mutants defective in spreading in cultured cells are compro-

mised for virulence in a mouse animal model [20,111–113].

Sites of bacterial spread in infected animals, as indicated by his-

tological studies, include the intestinal epithelium [114,115] and the liver [112]. Cell–cell spread not only facilitates colonization of

key host organs, but also contributes to infection of the fetus in

pregnant animals [116,117].

( a )

(i) linear actin filament

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016 branched actin filament

N-W A SP N-W A SP

A rp2/3 complex

A rp2/3 complex

(ii) N-W A SP (inactive)

WH 1 GBD

A C W

( b )

FH 1

WH 1

N-W A SP (active)

W C

P

A

PlP2, SH3 domains phosphorylation

Cdc 4 2 (Toca1 ,

Grb2, Nck)

4

Figure 2.

Actin polymerization mediated by the human Arp2/3 complex or formin proteins. (

a

) Arp2/3-dependent actin polymerization. (i) Cooperation of Arp2/3 and N-WASP to promote actin filament assembly. Arp2/3 is a seven-protein complex that stimulates the assembly of a new actin filament at the side of an existing

filament [50]. Arp2/3-mediated F-actin assembly requires the participation of NPFs such as N-WASP. N-WASP uses its WCA domain to stimulate Arp2/3-mediated

actin polymerization. The C and A regions of this domain bind and activate the Arp2/3 complex, whereas the W region delivers actin monomers to the Arp2/3 nucleation machinery. (ii) Regulation of N-WASP activity. N-WASP is subject to autoinhibition mediated by binding of its GTPase-binding domain (GBD) to the C and

A regions. The protein WIP stabilizes the inactive conformation of N-WASP. Autoinhibition of N-WASP is relieved by binding of the phospholipid phosphatidylinositol

4,5-bis phosphate (PIP2), the activated GTPase Cdc42 or SH3 domains of several signalling proteins. In addition, N-WASP can be activated by phosphorylation of a serine residue in its WCA domain. (

b

) Formin-mediated actin polymerization. Formin proteins function as dimers and use two domains to stimulate the assembly of

linear actin filaments [50,53]. The formin homology 2 (FH2) domain nucleates actin filaments, and the formin homology 1 (FH1) domain delivers profilin – actin

complexes to the filament’s barbed end. A, acidic region; C, connector region; CK2, casein kinase 2; SH3, Src Homology 3; W, WASP Homology 2 (WH2) domain; WIP,

WASP-interacting protein.

ABM, the first step in Listeria spread, was first described

nearly 25 years ago [8]. Since this discovery, seminal work

from several research groups has partly elucidated the biophysics of ABM and identified the bacterial and host factors that contribute to this process.

Listeria actively induces the polymerization of host actin filaments, a process that provides

the driving force for cytoplasmic movement of bacteria [118].

The bacterial factor responsible for ABM is a surface protein

called ActA [11,12]. Remarkably, ActA acts as a structural

and functional mimic of the eukaryotic NPF N-WASP

[1,2,51]. The amino-terminal domain of ActA contains

sequences with amino acid similarity to C and A regions of

N-WASP (figure 3

a

) [51,60,119]. This domain also has an

actin monomer binding sequence that is a functional equivalent

of the N-WASP W (WH2) region [119–121]. Like N-WASP, the

amino-terminal domain of ActA activates the Arp2/3 complex,

stimulating nucleation of branched actin filaments [60]. In

addition to this amino-terminal domain, a central proline-rich region of ActA also contributes to ABM by binding the host

protein VASP [122–126] One possible role of VASP is to recruit

profilin, which promotes addition of actin monomers to the

plus end of actin filaments [55,56,126,127].

It is noteworthy that ActA was the first protein, bacterial or eukaryotic, demonstrated to function as an NPF. In fact, work with ActA led to the discovery of the Arp2/3 complex as a machine that promotes the nucleation of actin filaments

[13,60]. After these initial studies with ActA, N-WASP and

other eukaryotic proteins were demonstrated to act as NPFs

for Arp2/3 [77–79,128,129]. Thus, studies with ActA serve as

a prime example of how microbial virulence proteins can be used as tools to understand fundamental aspects of eukaryotic cell biology.

In contrast to N-WASP, ActA lacks domains that mediate regulation by host GTPases. However, recent results indicate

( a ) ( b )

FH2L

A W C

P

PRD M phosphorylation

(CK2b

)

FH 1 L

W

W

W

FH 1 L

A T

FH2L

FH 1 L

W

W

W

FH 1 L

A T

Figure 3.

Bacterial surface proteins that stimulate actin polymerization.

(

a

) Structure of the

Listeria

protein ActA. ActA is a bacterial NPF with structural

and functional similarities to eukaryotic N-WASP [1,50]. An amino-terminal

domain in ActA has sequences with amino acid similarity to the W, C and

A regions in N-WASP. ActA’s central proline-rich domain (PRD) serves a function similar to that of the PRD in N-WASP. A carboxyl-terminal membrane anchoring domain (M) is responsible for association of ActA with the bacterial surface.

(

b

) Putative structure of the

Rickettsia

protein Sca2. Sca2 has regions with amino acid or secondary structural similarity to formin FH1 or FH2 domains,

respectively [35,36]. Here, these regions in Sca2 are referred to as ‘FH1-like

(FH1L)’ or ‘FH2-like’ (FH2L), respectively. Sca2 also has three WASP homology

2 (WH2) domains predicted to bind actin monomers. A carboxyl-terminal autotransporter (AT) domain in Sca2 is thought to anchor the protein to the outer

membrane of bacteria [36]. Because of the sequence or structural similarity of

Sca2 regions to domains in eukaryotic forms, Sca2 is depicted as a dimer.

However, the true oligomerization state of Sca2 is not known.

that ActA and N-WASP share a common regulatory mechanism involving serine phosphorylation in the vicinity of their C

regions [130] (figure 3

a ). The C regions of N-WASP and another eukaryotic NPF called WAVE2 are immediately adjacent to consensus phosphorylation sequences for the eukaryotic serine/ threonine kinase casein kinase 2 beta (CK2b

) [95]. Importantly,

CK2b -mediated phosphorylation of these sequences increases the affinity of N-WASP or WAVE2 for the Arp2/3 complex

and enhances Arp2/3-dependent actin polymerization [95].

Strikingly, the C region in ActA contains adjacent CK2 consensus phosphorylation sites similar to those in N-WASP and WAVE2

[130]. A variety of biochemical and genetic approaches were

used to demonstrate that CK2b -mediated phosphorylation of one of these sites in ActA occurs in vitro and in infected human cells. Moreover, mutational analysis of ActA indicates that phosphorylation is needed for efficient interaction with the host

Arp2/3 complex, bacterial ABM and full Listeria virulence in an animal model. These findings reveal that eukaryotic NPFs and the Listeria NPF ActA share a common post-translational regulatory mechanism. ActA can therefore be viewed not only as structurally and functionally mimicking eukaryotic NPFs, but also as exploiting the same host regulatory machinery. This

latter facet has been termed ‘regulatory mimicry’ [130]. Since

ActA and eukaryotic NPFs do not share extensive amino acid similarity except in their C regions and CK2b phosphorylation sites, it has been suggested that these proteins arose through con-

vergent evolution [130,131]. If so, then the shared regulation of

eukaryotic and microbial NPFs by CK2b would suggest that control of NPFs is critical, and a limited number of solutions exist for regulating C region activity.

3.2.2.

Shigella

Shigella flexneri is a Gram-negative bacterial pathogen that infects

cells of the intestinal epithelium, resulting in dysentery [5]. Intra-

cellular motility of Shigella was first described in the late 1960s

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016

[132], and movement was demonstrated to be actin-dependent

about 20 years later [26]. Based on the work in an animal

model, the ability of Shigella to undergo ABM and spread

between intestinal epithelial cells is crucial for disease [133].

ABM of Shigella is induced by the bacterial surface protein

IcsA, which is also known as VirG [26,27,29,30]. IcsA is unre-

lated in amino acid sequence to Listeria ActA, and these two bacterial proteins promote actin filament assembly through distinct mechanisms. Whereas ActA acts as a mimic of eukaryotic N-WASP, Shigella IcsA stimulates actin polymeriz-

ation by using host N-WASP [1,49,51]. In infected human

cells, N-WASP accumulates at the bacterial pole that produces

the F-actin tail [28,134]. Bacterial recruitment of host N-WASP

is due to IcsA, which uses an amino-terminal domain with glycine-rich repeats to bind directly to the human protein

[28,134]. Importantly, N-WASP is essential for ABM of

Shigella , as determined by experiments involving dominant negative N-WASP alleles or mouse cell lines deleted for the

N-WASP gene protein [28,135,136].

In vitro experiments with purified proteins demonstrate that IcsA is capable of activating N-WASP, resulting in Arp2/3-dependent actin

polymerization [134]. Interestingly, IcsA resembles the

eukaryotic GTPase Cdc42 in its ability to activate N-WASP, suggesting that the bacterial protein could be considered a functional mimic of Cdc42. In agreement with this idea,

ABM of Shigella

is independent of Cdc42 [137,138].

A large proportion of cellular N-WASP is complexed with

WIP, which stabilizes the autoinhibited form of N-WASP [80].

Cdc42-GTP alone is unable to activate N-WASP associated with WIP. Instead, activation requires the simultaneous

presence of Cdc42 and an additional protein called Toca-1 [94].

Interestingly, WIP is recruited to motile Shigella

[139], suggesting

that bacteria need to overcome WIP-mediated inhibition of

N-WASP in order to form F-actin comet tails. Consistent with this idea, recent results indicate a critical role for Toca-1 in

ABM of Shigella

[140]. Host Toca-1 associates with intracellular

bacteria immediately prior to actin-based movement. Importantly, recruitment of Toca-1 is independent of IcsA and is instead mediated by an unidentified Shigella factor that is injected into the host cell through a bacterial apparatus termed

a ‘type III secretion system’ [49]. Experiments involving RNAi-

mediated depletion of Toca-1 indicate a crucial role for this

human protein in bacterial-induced F-actin tail assembly [140].

Moreover, a constitutively activated derivative of N-WASP restores normal F-actin tail formation in cells depleted for

Toca-1. This latter result suggests that Toca-1 controls Shigella

ABM by contributing to N-WASP activation. Taken together

with previous studies [28,134], these recent findings indicate

that Shigella stimulates host Arp2/3-dependent actin polymerization by recruiting N-WASP via IcsA and exploiting Toca-1 to activate N-WASP.

5

3.2.3.

Rickettsia

The spotted fever group of Rickettsia cause severe systemic diseases characterized by infection of endothelial cells and

increased microvascular permeability [4]. ABM and cell–cell

spread of Rickettsia are thought to contribute to bacterial colonization of endothelial cells and resulting vascular dysfunction.

Interestingly, the F-actin tails of Rickettsia differ in structure from those of Listeria or Shigella . Whereas the latter bacteria have tails with a meshwork of branched actin filaments, Rickettsia

tails comprise parallel bundles of linear actin filaments [1,141].

The linear filaments in Rickettsia tails suggest that this bacterium might not use the host Arp2/3 complex to induce actin polymerization. Indeed, several reports indicate that inhibition of Arp2/3 fails to affect ABM of Rickettsia , in contrast to the situation observed with Listeria or Shigella

[142,143].

Recent results have shed light on how Rickettsia produces tails with linear actin filaments. The bacterial surface protein

Sca2 was shown to be required for ABM [36] and to nucleate

F-actin [35]. Interestingly, multiple lines of evidence indicate

that Sca2 may be a structural and functional mimic of eukaryotic formin proteins. First, Sca2 has domains with structures

and/or functions similar to regions in formins (figure 3 b

).

Specifically, the amino-terminal domain of Sca2 is predicted to share secondary structure similarity with the FH2 domains

of formins [35], and a central region in Sca2 has amino acid similarity to FH1 domains [36]. Sca2 also has three WASP

homology 2 (WH2) domains that are predicted to bind actin

monomers [35,36]. Second, the biochemical activities of Sca2 resemble those of formins [35]. Like formin proteins, Sca2 nucleates the assembly of linear actin filaments [35]. In addition, Sca2

promotes filament elongation by using profilin and inhibiting the activity of capping proteins. Importantly, Sca2 is needed

for virulence in an animal model [36], demonstrating that

ABM mediated by this bacterial protein contributes to disease.

It has been proposed that Rickettsia may have factors apart

from Sca2 that contribute to ABM [1]. The

Rickettsia protein

RickA has amino acid similarity to the WCA region of

N-WASP and is capable of stimulating Arp2/3-dependent actin polymerization in vitro

[37,38]. However, a major function for

RickA in ABM seems unlikely given the linear nature of actin filaments in Rickettsia

tails [141], the absence of Arp2/3 in these structures [141] and the lack of requirement for Arp2/3 in

Rickettsia

motility in human cells [142,143]. It is possible that the major

function of RickA is to induce actin cytoskeletal rearrangements involved in internalization of Rickettisa

into host cells [35].

4. Bacterial protrusion formation

While much progress has been made in dissecting mechanisms of bacterial ABM, considerably less is understood about the subsequent steps of membrane protrusion formation and engulfment. A key issue in the field has been whether pathogen-containing protrusions develop simply as a passive consequence of actin-based movement or whether instead these structures are actively controlled by bacterial

and/or host factors [43]. Although this area of research is

in its infancy, recent results suggest an active involvement of pathogen and host in the generation of protrusions

[44 –46]. Key findings with

Listeria and Shigella are described later. At present, studies on mechanisms of protrusion production by Rickettsia have not been reported.

4.1.

Listeria

Recent results have led to the identification of Listeria and human proteins that control the generation of bacterial protrusions in polarized human intestinal epithelial cells. After internalization into host cells, cytoplasmic Listeria secretes a

protein called InlC [144] that acts after F-actin tail assembly to

enhance protrusion formation and cell–cell spread [45]. InlC

promotes protrusions by physically interacting with and antagonizing the function of a human cytoplasmic protein called

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016

Tuba. Tuba is a large scaffolding protein with several functional domains, including a carboxyl-terminal Src Homology 3 (SH3) domain that associates with human N-WASP. Biochemical data demonstrate that InlC binds directly to the Tuba SH3 domain, thereby displacing N-WASP. Experiments with Listeria expressing a mutant InlC protein defective in binding Tuba indicate that the ability of InlC to displace N-WASP is critical for bacterial spread in cultured cells. Altogether, these findings demonstrate that Listeria enhances its spreading by disrupting complexes composed of host Tuba and N-WASP.

Studies on the role of InlC in spread have recently been

extended to an animal model [112]. This work took advan-

tage of a mutant InlC protein that folds normally but is compromised in binding to the Tuba SH3 domain.

Listeria expressing this mutant InlC protein has a virulence defect in intravenously inoculated mice that is similar to the defect of an inlC deletion strain. In addition, the Listeria mutant strain producing InlC defective in binding Tuba exhibits decreased cell– cell spread in the mouse liver. These studies support the idea that the ability of InlC to interact with host Tuba is important for Listeria virulence.

What are the normal functions of human Tuba and

N-WASP, and how does antagonism of these two host proteins by bacterial InlC enhance Listeria spread? In epithelial cells, Tuba and N-WASP act together to control morphology

of the apical junction complex [145]—a structure composed of tight junctions and adherens junctions [146]. Adherens

junctions promote cell –cell adhesion, whereas tight junctions act as a barrier to limit permeability to macromolecules and

ions [146,147]. Tight junctions also contribute to cell polarity

by establishing apical and basolateral plasma membrane domains. Importantly, RNAi-mediated depletion of Tuba or

N-WASP causes tight junctions to become slack, suggesting

a loss of cortical tension [45,145]. Thus, one of the normal

functions of Tuba/N-WASP complexes is in the maintenance of proper junctional structure, possibly by generating tension

through actin polymerization (figure 4 a

(i)). Interestingly, infection with Listeria or ectopic expression of InlC causes tight junctions to slacken, similar to the effects of Tuba or

N-WASP depletion [45] (figure 4

a (ii)). These findings indicate that InlC perturbs cell –cell junctions, probably through inhibition of Tuba and N-WASP. Altogether, the results suggest that Tuba/N-WASP complexes impose a potential barrier to bacterial spread by generating tension at cell junctions. This tension is expected to limit spreading by opposing the protrusive force of motile bacteria. By producing InlC, Listeria has evolved a mechanism to counteract the host machinery that normally generates cortical tension.

Apart from Tuba and N-WASP, another host factor that contributes to Listeria protrusion formation is the cytoskeletal

regulatory protein ezrin [44]. Ezrin is a member of the ERM

(ezrin/radixin/moesin) family of proteins [148]. ERM proteins

possess a carboxyl-terminal domain that interacts with F-actin

[149,150] and an amino terminal domain that binds to plasma membrane-associated proteins [151,152]. These domains

provide ERM proteins with the ability to link the actin cytoskeleton to the plasma membrane. ERM protein activity is subject to autoinhibition mediated by interaction of the amino- and

carboxyl-terminal regions [148]. One of the ways that autoinhi-

bition is relieved is through phosphorylation of key threonine residue (T567) in the carboxyl-terminal domain of ERM proteins

[153–155]. Importantly, the ERM protein ezrin localizes to

Listeria F-actin tails in protrusions, but not to tails in the main

6

( a )

(i)

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016

(ii) bTJ tTJ

= cortical tension

= Tuba

= N-W A SP

= lnlC

( b )

(i) (ii) bTJ tTJ clathrin epsin1 dynamin 2

Figure 4.

Spreading of

Listeria

and

Shigella

involves remodelling of host cell – cell junctions. (

a

) The formation of

Listeria

protrusions is enhanced by bacterialinduced alterations of tight junctions (TJs). (i) The human proteins Tuba and N-WASP form a complex that helps maintain linear TJs, probably by generating cortical

tension at the plasma membrane [45,145]. bTJ and tTJ denote ‘bicellular tight junction’ and ‘tricellular tight junction’, respectively, which are formed by the

intersection of two or three cells, respectively. (ii) In human cells infected with

Listeria

, the secreted bacterial protein InlC binds to host Tuba, thereby disrupting

Tuba – N-WASP complexes [45]. This inhibition of Tuba and N-WASP results in slack junctions, which are likely to reflect diminished cortical tension. The relief in

tension is thought to facilitate bacterial spread by removing an inward force at the host plasma membrane that would otherwise counteract the outward force exerted by motile bacteria. (

b

)

Shigella

protrusions are internalized through a host endocytic pathway at tTJs. (i) Live imaging studies indicate that

Shigella

contact

with tTJs leads to productive spreading [48]. (ii) tTJs affect the internalization of bacterial protrusions. Several human proteins known to promote clathrin-mediated

endocytosis (clathrin, epsin-1, dynamin 2) are needed for engulfment of

Shigella

protrusions, suggesting that bacteria may usurp a host endocytic pathway to facilitate their spread. Epsin-1 and clathrin are depicted as accumulating in coated pits in the host cell receiving the protrusion, and dynamin 2 is shown mediating scission of the protrusion. These activities of epsin-1, clathrin and dynamin 2 are speculative.

body of infected cells [44,141,156]. Inhibition of ERM proteins

through RNAi or expression of dominant negative ezrin alleles inhibits the formation of Listeria

protrusions [44]. Experiments

with an ezrin protein mutated in the T567 site indicate that activation of ERM proteins through phosphorylation is crucial for bacterial spread. Collectively, these results demonstrate an important role for host ERM proteins in the generation of Listeria protrusions. Interestingly, conditions that impair ERM protein function not only reduce the number of bacterial protrusions per host cell, but also alter the morphology of the few protrusions that are made. Compared with Listeria protrusions made under normal conditions, protrusions formed in cells with reduced ERM protein activity are shorter and wider, particularly in the region where the protrusion joins the main body of the cell. This aberrant morphology is consistent with the idea that ERM proteins may confer rigidity to Listeria protrusions by cross-linking F-actin tails to the host plasma membrane. This rigidity could contribute to cell–cell spread by allowing bacteria initiating protrusions to resist inward tension at cell junctions.

4.2.

Shigella

Specific bacterial factors involved in the formation of Shigella containing protrusions have not been identified. Interestingly, the same bacterial type III secretion system that promotes internalization of Shigella into host cells is also

needed for cell –cell spread [157]. As mentioned in §3.2.2,

one of the ways that this type III secretion system controls spreading is by stimulating F-actin tail assembly through

recruitment of Toca-1 [140]. Whether the secretion system

also acts after ABM to directly affect protrusion formation is not known. It would be interesting to screen known bacterial substrates of this secretion system for roles in the generation of bacterial protrusions.

Host factors that promote the formation of Shigellacontaining protrusions include formin proteins, the actin-dependent motor protein myosin X and the cell–cell adhesion molecule

E-cadherin. The role of formins appears to be in stimulating the assembly of actin filaments in comet tails in protrusions

[158]. Thus, while

Shigella F-actin tail formation in the main body of the cell requires host N-WASP and the Arp2/3 complex, actin polymerization in protrusions is thought to switch to a formin-mediated pathway. One of the key findings in this study is that the diaphanous formin Dia1 localizes to

F-actin tails in the protrusions, but not to those in the cell

body [158]. In addition, inhibition of Dia1 or Dia1 formins

through RNAi or dominant negative approaches reduces the frequency of protrusion formation by Shigella . It was proposed that the switch to formin-mediated actin polymerization facilitates bacterial protrusion generation by re-organizing the dense cortical actin network that underlies the plasma mem-

brane [158]. If not remodelled, this network would be

expected to limit contact of motile bacteria with the plasma

7

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016 membrane. A second role of formin-mediated actin polymerization in comet tails may be to generate force necessary to deform the plasma membrane into protrusions.

Myosin X is a ubiquitously expressed unconventional myosin involved in filopodia formation and phagocytosis

[159]. A recent report demonstrates an important role for

this myosin in the formation of Shigella protrusions leading

to cell –cell spread [46]. Myosin X is recruited to

Shigella

F-actin tails in protrusions [46]. RNAi-mediated depletion

of myosin X reduces the length of Shigella protrusions, without affecting the number of protrusions produced per infected cell. These findings indicate that myosin X is dispensable for the initiation of protrusions, but required for their extension. Myosin X has several domains, including a head domain with ATP-dependent motor activity and a Pleckstrin homology (PH) domain that associates with the inner face

of the plasma membrane [159]. The ability of myosin X

to extend protrusions requires both of these domains [46].

Interestingly, time-lapse microscopy indicates that myosin X localization to protrusions is dynamic, with myosin clusters in F-actin tails cycling towards or away from protrusion tips. This cycling may reflect the movement of myosin motors along the actin filaments in tails. Based on these data, a model was proposed whereby the head domain of myosin X interacts with bacterial F-actin tails in an emerging protrusion and the PH domain associates with the plasma

membrane [46]. According to this model, ATP-dependent

motor activity causes myosin X to ‘walk’ along F-actin filaments in the tail, in the direction of filament plus ends.

Since the myosin is anchored to the host plasma membrane, its directional movement along F-actin tails stimulates translocation of membrane towards the protrusion tip, driving protrusion growth. While attractive, this model may at first glance seem at odds with the observation in time-lapse microscopy that myosin X sometimes moves towards the

base of protrusions [46]. Perhaps this movement is a recycling

process that allows the same clusters of myosin X to be used multiple times for protrusion growth.

Interestingly, in addition to decreasing Shigella protrusion length, depletion of myosin X increases width in the region

where protrusions intersect the main body of the cell [46].

This effect on Shigella protrusion width shows striking similarity to the role of ERM proteins in controlling the width of Listeria protrusions. These findings suggest that a second function of myosin X may be to confer rigidity to protrusions by linking F-actin tails to the plasma membrane.

An early study demonstrated an important role for host

E-cadherin in the generation of Shigella protrusions of proper structure and also in the internalization of these pro-

trusions by neighbouring host cells (see §5) [160]. This

study used fibroblast cell lines lacking or expressing chicken

E-cadherin. Scanning and transmission electron microscopy analysis indicated that protrusions made in cells lacking

E-cadherin were flaccid, lacking tight association of bacteria and F-actin tails with the host plasma membrane. In addition, immunofluorescence studies revealed the presence of adherens junction components in F-actin tails in protrusions.

Interestingly, these comet tails appear to intersect cell –cell junctions, suggesting that protrusions emanate from adherens junctions. Collectively, the results in this study indicate the E-cadherin-mediated cell– cell adhesion is essential for

Shigella spread, and that bacterial protrusion formation may involve targeting of adherens junctions.

5. Internalization of protrusions

An important question is whether Listeria , Shigella or Rickettsia produce factors that stimulate the ability of host cells to engulf protrusions. Thus far, such microbial factors have not been identified. In recent years, some progress has been made in identifying host proteins specifically involved in the internalization of Listeria or Shigella protrusions. These findings are discussed later.

8

5.1.

Listeria

An early study proposed that engulfment of Listeria protrusions might occur through a host-driven process that is

active even in the absence of microbial infection [161]. This

proposal was based on the observation that vesicles produced by labelled uninfected epithelial cells were internalized by

neighbouring cells [161]. The overall size and shape of these

vesicles resembled those observed during cell–cell spread of

Listeria . The process of membrane internalization between neighbouring cells was termed ‘paracytophagy’.

Despite these interesting early observations, progress in identifying human factors mediating engulfment of Listeria protrusions has been slow. So far, only one such factor has been found—human casein kinase 1 alpha (CK1a

) [47]. Identi-

fication of CK1a was accomplished through a high-throughput

RNAi-based screen in the human cell line HeLa. This screen involved assessing the roles of 779 known human kinases in cell–cell spread of Listeria . After performing experiments to exclude off-target effects of siRNAs, two of the original approximately 800 kinases investigated were found to have bona fide roles in spreading. These kinases were CK1a and casein kinase 2 beta (CK2b ). As described in §3.2.1, CK2b controls the actin polymerization step of Listeria spread through phosphorylation of ActA. Further analysis of CK1a indicated that this kinase is dispensable for F-actin comet tail formation and the formation of protrusions, but is needed for the resolution of protrusions into vacuoles containing Listeria . Interestingly,

CK1a controls bacterial spread by acting in the infected host cell that donates the protrusion, not in the neighbouring cell that receives the protrusion. How CK1a accomplishes protrusion resolution is not known, and solving this mystery may require additional RNAi-based screens of known casein kinase

1 substrates. It is possible that CK1a phosphorylates a human protein involved in plasma membrane scission. This idea is discussed further in §6.

5.2.

Shigella

As mentioned in §4.2, E-cadherin is required not only for the formation of Shigella protrusions of normal morphology, but also for the internalization of these structures by neighbour-

ing cells [161]. It is not currently known if there is a direct

role for E-cadherin or other adherens junction components in uptake of protrusions or whether the internalization defect in E-cadherin deficient cells is a secondary consequence of the aberrant protrusions formed. In addition to adherens junctions, gap junctions have also been implicated in the intercellular dissemination of Shigella

[162]. The gap

junction component connexin 26 was found to enhance the spreading of Shigella through a mechanism involving the release of ATP into the external medium. The specific stage

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016 of spreading affected by connexin 26 (e.g. F-actin tail formation, protrusion formation, protrusion engulfment) was not addressed. Altogether, these findings highlight the importance of cell –cell junctions in Shigella spread, a general feature now known to be shared with Listeria

[45].

A recent study has investigated the role of host junctions in

Shigella spread in further detail. Using time-lapse microscopy,

Shigella was found to spread predominantly through contact

with tricellular tight junctions (tTJs) [48], areas in an epithe-

lial cell monolayer formed by the intersection of three cells

(figure 4 b

(i)) [163]. By comparison, bicellular tight junctions

(bTJs) are defined by the intersection of two cells. Despite the fact that tTJs occupy a very small area relative to bTJs, approximately 80% of cell–cell spread events with Shigella

occur through tTJs [48]. These findings demonstrate that tTJs

are membrane contact sites allowing highly efficient spread of Shigella . About 40–50% of Listeria spreading events occur at tTJs, indicating a lesser but still significant role for tTJs in

dissemination of this pathogen [48]. An important component

of tTJs is the protein tricellulin [163,164]. Tricellulin localizes

predominantly to tTJs and is also found in lesser abundance at bTJs. Importantly, RNAi-mediated depletion of tricellulin impairs cell–cell spread of Shigella , indicating a functional

role for tricellulin in this process [48]. In addition, spreading

of Shigella through tTJs requires several human proteins with known roles in endocytosis, including epsin-1, clathrin and

dynamin 2 [48]. Interestingly, depletion of tricellulin or these

endocytic proteins does not affect the number of Shigella protrusions formed per cell, suggesting that these proteins

mediate protrusion engulfment (figure 4 b

(ii)). In support of this idea, fluorescence microscopy analysis of live infected human cells reveals recruitment of endocytic proteins in cells

internalizing protrusions [48].

Why does Shigella spread occur predominantly at tTJs? One possibility is that tTJs are regions of high endocytic activity, a property that would explain the apparent preference for internalization of protrusions in these regions. This idea is pure speculation, since to the best of my knowledge endocytosis at tTJs has never been investigated. It is worth remarking that tTJs create a channel or ‘central tube’ approximately 10 nm

wide in the epithelial monolayer [163]. The plasma membrane

surrounding this channel might have a protein and/or lipid composition distinct from that in bTJs—differences that could potentially affect many processes, including endocytosis.

6. Conclusions and outstanding questions

Since the first description of cell–cell spread nearly 25 years ago, the molecular basis of F-actin assembly by Listeria and Shigella has been extensively characterized to the point where it is possible to reconstitute ABM with purified bacterial and host

proteins [165]. Substantial progress has recently been made on

the mechanism of ABM by Rickettsia

[35,36,143]. Studies on

Listeria , Shigella and Rickettsia have revealed that ABM occurs through exploitation of host proteins (e.g. N-WASP and Arp2/

3) and/or the action of bacterial proteins (ActA and Sca2) that structurally and functionally mimic eukaryotic NPFs or

formins [11–13,26–30,35,36,134]. Interestingly,

Listeria or

Shigella subvert host regulatory mechanisms involving serine phosphorylation or the N-WASP activator Toca-1 to promote

F-actin assembly [130,140]. The molecular bases of bacterial pro-

trusion formation and engulfment are less well understood than

ABM and have been the subject of several recent studies. Findings with Listeria and Shigella indicate that bacteria target host cell–cell junctions to facilitate the generation or the internaliz-

ation of protrusions [45,48]. In addition, the formation of

Shigella

protrusions involves myosin motor activity [46] and a

switch from Arp2/3- to formin-mediated actin assembly [158].

These recent findings on bacterial protrusions prompt a variety of important questions to address in future work. Some of these questions are outlined as follows.

9

— Does protrusion formation by Listeria involve a transition to formin-mediated actin polymerization ? Interestingly, Listeria

F-actin tails in protrusions consist predominantly of linear filaments, in contrast to the branched F-actin net-

work in tails in the host cell body [156]. The linear

nature of filaments in protrusions raises the possibility that Listeria , like Shigella

[158], undergoes a switch to

host formin-mediated actin polymerization.

— What host cell process is antagonized by Listeria to perturb apical junctions?

The Listeria protein InlC alters the structure of apical junctions by inhibiting human Tuba and

N-WASP [45]. Elucidating the mechanism by which InlC

affects junctions will first require understanding how

Tuba and N-WASP normally control junctional structure.

These two human proteins could directly affect junctions through the generation of F-actin involved in actomyosin-

mediated tension [166]. Alternatively, Tuba and N-WASP

might indirectly impact junction morphology through control of membrane trafficking pathways. For example,

N-WASP promotes endocytosis [50,57] and vesicular

trafficking at the Golgi apparatus [167– 169]. Tuba is localized at sites of endocytosis and the Golgi [170,171],

raising the possibility that this protein aids N-WASP in membrane trafficking.

— Does cell –cell spread of Shigella and/or Rickettsia involve relief of tension at cell –cell junctions?

Similar to Listeria ,

Shigella infects polarized cells of the intestinal epithelium

[5]. These epithelial cells contain AJs and TJs subject to

cortical tension [147,166].

Rickettsia spreads in human

endothelial cells [4], another polarized cell type with AJs

and TJs [147]. An interesting question is how

Shigella and/or Rickettsia cope with cortical tension at junctions.

Have these pathogens evolved strategies similar to Listeria ’s ability to dissipate tension?

— How does host membrane remodelling occur at bacterial protrusions?

Protrusions have negative curvature at their tip and

positive curvature at their base (figure 5

a (i)). How is the plasma membrane of the host cell reshaped to form bacterial protrusions? Remodelling of eukaryotic membranes into curved shapes is accomplished by a variety of membrane-bending proteins of the BAR domain superfamily

[172]. Some BAR domain proteins have convex mem-

brane-binding surfaces, allowing them to induce negative membrane curvature. Other BAR domain proteins use concave surfaces to impart positive curvature to membranes.

Importantly, several BAR proteins with convex or concave membrane-binding surfaces promote the formation of

filopodia in mammalian cells [57]. Filopodia are plasma

membrane projections that superficially resemble bacterial protrusions, but are of smaller diameter. In future work, it will be interesting to investigate the role of these BAR domain proteins in protrusion formation by Listeria ,

Shigella or Rickettsia .

(

( a )

(i)

B A R,

F-B A R?

I-B A R?

b )

2

1

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016

(ii) exocytosis?

myosin X

Figure 5.

Potential mechanisms controlling the formation and engulfment of bacterial protrusions. (

a

) Formation of protrusions. (i) Remodelling of the host plasma membrane during the initiation of protrusion formation requires generation of negative and positive curvature at the protrusion tip and base, respectively. Several classes of human BAR proteins capable of producing

negative or positive curvature exist [57,172]. I-BAR proteins have convex

plasma membrane-binding domains and may help to remodel membrane at protrusion tips. BAR and F-BAR proteins have concave membrane-binding domains and could act the base of protrusions. (ii) Extension of protrusions.

The motor protein myosin X promotes the growth of

Shigella

protrusions, possibly by transporting host plasma membrane towards the protrusion tip

[46]. An unresolved question is whether extension of bacterial protrusions

also requires the localized delivery of new host membrane through exocytosis.

(

b

) Engulfment of protrusions. Internalization and conversion of a protrusion to a double membranous vacuole requires the scission and rejoining of plasma membrane in the human cell donating the protrusion (i) and also in the cell receiving the protrusion (ii). Host or bacterial factors responsible for these two scission/rejoining events are yet to be identified.

— How do bacterial protrusions extend?

Experiments with Shigella suggest that myosin X may promote the growth of protrusions by transporting host plasma membrane towards the

protrusion tip [46]. Could protrusion extension also involve

the local insertion of new host membrane delivered from intracellular compartments? Apical junctions, structures exploited by Listeria and Shigella for spread, are active sites

of exocytosis [146,173–175]. Interestingly, these junctions are located in close proximity to the exocyst [176]—a multi-

component machinery that promotes tethering and fusion

of intracellular vesicles with the plasma membrane [177].

An intriguing idea is that bacterial pathogens might exploit host exocytic activity at junctions to provide membrane

needed for protrusion growth (figure 5

a (ii)).

— Does internalization of protrusions involve exploitation of host endocytic pathways that normally control junctional integrity?

Apical junctions are sites of endocytosis, as well as exocyto-

sis [146,173,174]. Constitutive endocytosis of TJ and AJ

components is thought to be involved in epithelial tissue

homeostasis [146]. An important question is whether

bacterial pathogens hijack junctional endocytic routes to allow internalization of protrusions. In this regard, it would be interesting to determine the extent to which the endocytic pathway that promotes engulfment of Shigella

protrusions [48] also affects internalization of TJ and AJ

components in healthy epithelia. In addition, it will be important to assess the roles of known host endocytic proteins in engulfment of Listeria and Rickettsia protrusions.

— How are protrusions converted to double membranous vacuoles?

Bacterial protrusions contain two host plasma membranes

(figure 5 b

). One of these membranes originates from the human cell receiving the protrusion, and the other is provided by the cell donating the protrusion. Resolution of an internalized protrusion into the vacuole requires scission of both host-derived membranes followed by rejoining

(fusion) of these membranes. How scission and rejoining at different membranes are accomplished and coordinated is not understood. A candidate for a human protein mediating scission/rejoining in the cell receiving the protrusion is the GTPase dynamin 2. This GTPase promotes

scission during clathrin-mediated endocytosis [178] and is

also required for uptake of Shigella

protrusions [48]. It

seems unlikely, however, that dynamin 2 is involved in scission of the membrane from the cell donating the protrusion.

Dynamin proteins have never been observed to act within membrane tubules, but only on the outside of these tubules.

On the other hand, the mammalian endosomal complex required for sorting (ESCRT) stimulates membrane scission/rejoining from inside tubules to promote multivesicular body formation, shedding of membrane blebs

and virus budding [179]. It is possible that the ESCRT path-

way helps resolve bacterial protrusions by acting in the host cell donating the protrusion. As previously mentioned, human CK-1 a is needed for resolution of Listeria protrusions

[47], and this kinase could potentially regulate ESCRT or

other host scission machinery. Alternatively, unidentified bacterial factors might exert membrane scission/rejoining activity in the protrusion-donating cell.

10

Future work will undoubtedly answer some of the key questions described above. It is likely that high-throughput RNAi-

based screens [47] will play an increasingly important role in

the identification of host factors involved in cell–cell spread.

The choice of mammalian cell line used in such screens, and in the study of bacterial spread in general, is likely to be important. Given the role of cell–cell junctions in controlling dissemination of Listeria and Shigella , cell lines that have the capacity to develop into polarized monolayers with tight barriers are probably better models for post-ABM steps in spreading than cells that lack these characteristics. Whenever possible, it will also be important to confirm results obtained in tissue culture studies with experiments in animal models.

Classic histological studies have proved useful for assessing

bacterial spread at a few defined time points [112]. Real-time

imaging approaches with live animals [180] are expected to

allow more extensive analysis of spreading dynamics in vivo .

Future work with Listeria , Shigella and Rickettsia will not only help elucidate disease mechanisms, but should also contribute to a better understanding of important aspects of eukaryotic cell biology, including actin assembly, regulation of cell–cell junctions and membrane remodelling.

Acknowledgements.

I thank John Brumell, Suzanne Osborne, Antonella

Gianfelice and Luciano Rigano for comments on the manuscript.

Funding statement.

Work in my laboratory is supported by grants from the National Institutes of Health (R01AI085072), Marsden Fund of the Royal Society of New Zealand (UOO1003) and the Lottery

Health Fund of New Zealand.

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016

References

1.

Gouin E, Welch MD, Cossart P. 2005 Actin-based motility of intracellular pathogens.

Curr. Opin.

Microbiol.

8 , 35–45. (doi:10.1016/j.mib.2004.12.013)

2.

Haglund CM, Welch MD. 2011 Pathogens and polymers: microbe – host interactions illuminate the cytoskeleton.

J. Cell Biol.

195 , 7 – 17. (doi:10.1083/ jcb.201103148)

3.

Vazquez-Boland JA, Kuhn M, Berche P, Chakraborty

T, Dominguez-Bernal G, Goebel W, Gonzalez-Zorn B,

Wehland J, Kreft J. 2001 Listeria pathogenesis and molecular virulence determinants.

Clin. Microbiol.

Rev.

14 , 584 – 640. (doi:10.1128/CMR.14.3.

584-640.2001)

4.

Walker DH, Ismail N. 2008 Emerging and reemerging rickettsioses: endothelial cell infection and early disease events.

Nat. Rev. Microbiol.

6 ,

375 – 386. (doi:10.1038/nrmicro1866)

5.

Mavris M, Sansonetti PJ. 2004 Microbial-gut interactions in health and disease: epithelial cell responses.

Best Pract. Res. Clin. Gastroenterol.

18 ,

373 – 386. (doi:10.1016/j.bpg.2003.10.007)

6.

Ray K, Marteyn B, Sansonetti PJ, Tang CM. 2009 Life on the inside: the intracellular lifestyle of cytosolic bacteria.

Nat. Rev. Microbiol.

7 , 333 – 340. (doi:10.

1038/nrmicro2112)

7.

Stavru F, Archambaud C, Cossart P. 2011 Cell biology and immunology of Listeria monocytogenes infections: novel insights.

Immunol. Rev.

240 ,

160 – 184. (doi:10.1111/j.1600-065X.2010.00993.x)

8.

Tilney LG, Portnoy DA. 1989 Actin filaments and the growth, movement, and spread of the intracellular bacterial parasite Listeria monocytogenes .

J. Cell Biol.

109 , 1597 – 1608. (doi:10.1083/jcb.109.4.1597)

9.

Gaillard JL, Berche P, Frehel C, Gouin E, Cossart P.

1991 Entry of L. monocytogenes into cells is mediated by internalin, a repeat protein reminiscent of surface antigens from Gram-positive cocci.

Cell 65 , 1127 – 1141. (doi:10.1016/0092-8674(91)

90009-N)

10. Dramsi S, Biswas I, Maguin E, Braun L, Mastroeni P,

Cossart P. 1995 Entry of Listeria monocytogenes into hepatocytes requires expression of InIB, a surface protein of the internalin multigene family.

Mol.

Microbiol.

16 , 251 – 261. (doi:10.1111/j.1365-2958.

1995.tb02297.x)

11. Kocks C, Gouin E, Tabouret M, Berche P, Ohayon H,

Cossart P. 1992 Listeria monocytogenes -induced actin assembly requires the actA gene product, a surface protein.

Cell 68 , 521 – 531. (doi:10.1016/

0092-8674(92)90188-I)

12. Domann E, Wehland J, Rohde M, Pistor S, Hartl M,

Goebel W, Leimeister-Wa¨chter M, Wuenscher M,

Chakraborty T. 1992 A novel bacterial gene in Listeria monocytogenes required for host cell microfilament interaction with homology to the proline-rich region of vinculin.

EMBO J.

11 , 1981–1990.

13. Welch MD, Iwamatsu A, Mitchison TJ. 1997 Actin polymerization is induced by Arp2/3 protein complex at the surface of Listeria monocytogenes .

Nature 385 , 265 – 269. (doi:10.1038/385265a0)

14. Mengaud J, Ohayon H, Gounon P, Mege RM, Cossart

P. 1996 E-Cadherin is the receptor for internalin, a surface protein required for entry of L.

monocytogenes into epithelial cells.

Cell 84 ,

923 – 932. (doi:10.1016/S0092-8674(00)81070-3)

15. Shen Y, Naujokas M, Park M, Ireton K. 2000 InlBdependent internalization of Listeria is mediated by the Met receptor tyrosine kinase.

Cell 103 ,

501 – 510. (doi:10.1016/S0092-8674(00)00141-0)

16. Ireton K, Payrastre B, Chap H, Ogawa W, Sakaue H,

Kasuga M, Cossart P. 1996 A role for phosphoinositide 3-kinase in bacterial invasion.

Science 274 , 780 – 782. (doi:10.1126/science.274.

5288.780)

17. Sun H, Shen Y, Dokainish H, Holgado-Madruga M,

Wong A, Ireton K. 2005 Host adaptor proteins Gab1 and CrkII promote InlB-dependent entry of Listeria monocytogenes .

Cell. Microbiol.

7 , 443 – 457.

(doi:10.1111/j.1462-5822.2004.00475.x)

18. Veiga E, Cossart P. 2005 Listeria hijacks the clathrindependent endocytic machinery to invade mammalian cells.

Nat. Cell Biol.

7 , 894 – 900.

(doi:10.1038/ncb1292)

19. Geoffroy C, Gaillard JL, Alouf JE, Berche P. 1987

Purification, characterization and toxicity of the sulfhydryl-activated hemolysin listeriolysin O from

Listeria monocytogenes .

Infect. Immun.

55 ,

1641 – 1646.

20. Portnoy DA, Jacks PS, Hinrichs DJ. 1988 Role of hemolysin for the intracellular growth of Listeria monocytogenes .

J. Exp. Med.

167 , 1459 – 1471.

(doi:10.1084/jem.167.4.1459)

21. Vicente MF, Baquero F, Perez-Diaz C. 1985 Cloning and expression of the Listeria monocytogenes haemolysin in Escherichia coli .

FEMS Microbiol. Lett.

30 , 77 – 79. (doi:10.1111/j.1574-6968.1985.

tb00988.x)

22. Vazquez-Boland J-A, Kocks C, Dramsi S, Ohayon H,

Geoffroy C, Mengaud J, Cossart P. 1992 Nucleotide sequence of the lecithinase operon of Listeria monocytogenes and possible role of lecithinase in cell-to-cell spread.

Infect. Immun.

60 , 219 – 230.

23. Camilli A, Goldfine H, Portnoy DA. 1991 Listeria monocytogenes mutants lacking phosphatidylspecific phospholipase C are avirulent.

J. Exp. Med.

173 , 751 – 754. (doi:10.1084/jem.173.3.751)

24. Leimeister-Wa¨chter M, Domann E, Chakraborty T.

1991 Detection of a gene encoding a phosphatidylinositol-specific phospholipase C that is coordinately expressed with listeriolysin in Listeria monocytogenes .

Mol. Microbiol.

5 , 361 – 366.

(doi:10.1111/j.1365-2958.1991.tb02117.x)

25. Mengaud J, Braun-Breton C, Cossart P. 1991

Identification of a phosphatidylinositol-specific phospholipase C in Listeria monocytogenes : a novel type of virulence factor?.

Mol. Microbiol.

5 , 367 –

372. (doi:10.1111/j.1365-2958.1991.tb02118.x)

26. Bernardini ML, Mounier J, D’Hauteville H, Coquis-

Rondon M, Sansonetti PJ. 1989 Identification of icsA , a plasmid locus of Shigella flexneri that governs bacterial intra- and intercellular spread through interaction with F-actin.

Proc. Natl Acad.

Sci. USA 86 , 3867 – 3871. (doi:10.1073/pnas.

86.10.3867)

27. Lett M-C, Sasakawa C, Okada N, Sakai T, Makino S,

Yamada M, Komatsu K, Yoshikawa M. 1989 virG , a plasmid-coded virulence gene of Shigella flexneri :

Identification of the virG protein and determination of the complete coding sequence.

J. Bacteriol.

171 ,

353 – 359.

28. Suzuki T, Miki H, Takenawa T, Sasakawa C. 1998

Neural Wiskott – Aldrich syndrome protein is implicated in the actin-based motility of Shigella flexneri .

EMBO J.

17 , 2767 – 2776. (doi:10.1093/ emboj/17.10.2767)

29. Pal T, Newland JW, Tall BD, Formal SB, Hale TL.

1989 Intracellular spread of Shigella flexneri associated with the kcpA locus and a 140-kilodalton protein.

Infect. Immun.

57 , 477 – 486.

30. Makino S, Sasakawa C, Kamata K, Kurata T,

Yoshikawa M. 1986 A genetic determinant required for continuous reinfection of adjacent cells on large plasmid in S. flexneri 2a.

Cell 46 , 551 – 555. (doi:10.

1016/0092-8674(86)90880-9)

31. High N, Mounier J, Pre´vost MC, Sansonetti PJ. 1992

IpaB of Shigella flexneri causes entry into epithelial cells and escape from the phagocytic vacuole.

EMBO

J.

11 , 1991 – 1999.

32. Menard R, Sansonetti PJ, Parsot C. 1993 Nonpolar mutagenesis of the ipa genes defines IpaB, IpaC, and IpaD as effectors of Shigella flexneri entry into epithelial cells.

J. Bacteriol.

175 , 5899 – 5906.

33. Sasakawa C, Makino S, Kamata K, Yoshikawa M.

1986 Isolation, characterization, and mapping of

Tn5 insertions into the 140-megadalton invasion plasmid defective in the mouse Sereny test in

Shigella flexner i 2a.

Infect. Immun.

54 , 32 – 36.

34. Chan YGY, Riley SP, Martinez JJ. 2010 Adherence to and invasion of host cells by spotted fever group

Rickettsia species.

Front. Microbiol.

1 , 139. (doi:10.

3389/fmicb.2010.00139)

35. Haglund CM, Choe JE, Skau CT, Kovar DR, Welch

MD. 2010 Rickettsia Sca2 is a bacterial formin-like mediator of actin-based motility.

Nat. Cell Biol.

12 ,

1057 – 1063. (doi:10.1038/ncb2109)

36. Kleba B, Clark TR, Lutter EI, Ellison DW, Hackstadt T.

2010 Disruption of the Rickettsia rickettsii Sca2 autotransporter inhibits actin-based motility.

Infect. Immun.

78 , 2240 – 2247. (doi:10.1128/IAI.

00100-10)

37. Gouin E, Egile C, Dehoux P, Villiers V, Adams J,

Gertler FB, Li R, Cossart P. 2004 The RickA protein of

Rickettsia conorii activates the Arp2/3 complex.

Nature 427 , 29. (doi:10.1038/nature02318)

38. Jeng RL, Goley ED, D’Alession JA, Chaga OY, Svitkina

TM, Borisy GG, Heinzen RA, Welch MD. 2004 A

Rickettsia WASP-like protein activates the Arp2/3 complex and mediates actin-based motility.

Cell.

Microbiol.

6 , 761 – 769. (doi:10.1111/j.1462-5822.

2004.00402.x)

11

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016

39. Martinez JJ, Seveau S, Veiga E, Matsuyama S,

Cossart P. 2005 Ku70, a component of

DNAdependent protein kinase, is a mammalian receptor for Rickettsia conorii .

Cell 123 , 1013 – 1023.

(doi:10.1016/j.cell.2005.08.046)

40. Chan YG, Cardwell MM, Hermanas TM, Uchiyama T,

Martinez JJ. 2009 Rickettsial outer-membrane protein B (rOmpB) mediates bacterial invasion through Ku70 in an actin, c-Cbl, clathrin and caveolin 2-dependent manner.

Cell. Microbiol.

11 ,

629 – 644. (doi:10.1111/j.1462-5822.2008.01279.x)

41. Riley SP, Goh KC, Hermanas TM, Cardwell MM, Chan

YG, Martinez JJ. 2010 The Rickettsia conorii autotransporter protein Sca1 promotes adherence to nonphagocytic mammalian cells.

Infect. Immun.

78 ,

1895 – 1904. (doi:10.1128/IAI.01165-09)

42. Cardwell MM, Martinez JJ. 2009 The Sca2 autotransporter protein from Rickettsia conorii is sufficient to mediate adherence to and invasion of cultured mammalian cells.

Infect. Immun.

77 ,

5272 – 5280. (doi:10.1128/IAI.01165-09)

43. Monack DM, Theriot JA. 2001 Actin-based motility is sufficient for bacterial membrane protrusion formation and host cell uptake.

Cell. Microbiol.

3 ,

633 – 647. (doi:10.1046/j.1462-5822.2001.00143.x)

44. Pust S, Morrison H, Wehland J, Sechi AS, Herrlich P.

2005 Listeria monocytogenes exploits ERM protein functions to efficiently spread from cell to cell.

EMBO. J.

24 , 1287 – 1300. (doi:10.1038/sj.

emboj.7600595)

45. Rajabian T, Gavicherla B, Heisig M, Muller-Altrock S,

Goebel W, Gray-Owen SD, Ireton K. 2009 The bacterial virulence factor InlC perturbs apical cell junctions and promotes cell-to-cell spread of

Listeria .

Nat. Cell Biol.

11 , 1212 – 1218. (doi:10.

1038/ncb1964)

46. Bishai EA, Sidhu GS, Li W, Dhillon J, Bohil AB,

Cheney RE, Hartwig JH, Southwick FS. 2013 Myosin-

X facilitates Shigella -induced membrane protrusions and cell-to-cell spread.

Cell. Microbiol.

15 ,

353 – 367. (doi:10.1111/cmi.12051)

47. Chong R, Squires R, Swiss R, Agaisse H. 2011 RNAi screen reveals host cell kinases specifically involved in Listeria monocytogenes spread from cell to cell.

PLoS ONE 6 , e23399. (doi:10.1371/journal.pone.

0023399)

48. Fukumatsu M, Ogawa MFA, Suzuki M, Nakayama K,

Shimizu S, Kim M, Mimuro H, Sasakawa C. 2012

Shigella targets epithelial tricellular junctions and uses a noncanonical clathrin-dependent endocytic pathway to spread between cells.

Cell Host Microbe

11 , 325 – 336. (doi:10.1016/j.chom.2012.03.001)

49. Ogawa M, Handa Y, Ashida H, Suzuki M,

Sasakawa C. 2008 The versatility of Shigella effectors.

Nat. Rev. Microbiol.

6 , 11 – 16.

(doi:10.1038/nrmicro1814)

50. Campellone KG, Welch MD. 2010 A nucleator arms race: cellular control of actin assembly.

Nat. Rev.

Mol. Cell Biol.

11 , 237 – 251. (doi:10.1038/nrm2867)

51. Goldberg MB. 2001 Actin-based motility of intracellular microbial pathogens.

Microbiol. Mol.

Biol. Rev.

65 , 595 – 626. (doi:10.1128/MMBR.65.4.

595-626.2001)

52. Firat-Karalar EN, Welch MD. 2011 New mechanisms and functions and actin nucleation.

Curr. Opin. Cell

Biol.

23 , 4 – 13. (doi:10.1016/j.ceb.2010.10.007)

53. Chesarone MA, DuPage AG, Goode BL. 2010

Unleashing formins to remodel the actin and microtubule cytoskeletons.

Nat. Rev. Mol. Cell Biol.

11 , 62 – 74. (doi:10.1038/nrm2816)

54. Dominguez R, Holmes KC. 2011 Actin structure and function.

Ann. Rev. Biophys.

40 , 169 – 186. (doi:10.

1146/annurev-biophys-042910-155359)

55. Goldschmidt-Clermont PJ, Machesky LM, Doberstein

SK, Pollard TD. 1991 Mechanism of the interaction of human platelet profilin with actin.

J. Cell Biol.

113 , 1081 – 1089. (doi:10.1083/jcb.113.5.1081)

56. Pantaloni D, Carlier MF. 1993 How profilin promotes actin filament assembly in the presence of thymosin beta 4.

Cell 75 , 1007 – 1014. (doi:10.1016/0092-

8674(93)90544-Z)

57. Suetsugu S, Gautreau A. 2012 Synergistic BAR – NPF interactions in actin-driven membrane remodeling.

Trends Cell Biol.

22 , 141 – 150. (doi:10.1016/j.tcb.

2012.01.001)

58. Nightingale TD, Cutler DF, Cramer LP. 2012 Actin coats and rings promote regulated exocytosis.

Trends Cell Biol.

22 , 329 – 337. (doi:10.1016/j.tcb.

2012.03.003)

59. Scott CC, Dobson W, Botelho RJ, Coady-Osberg N,

Chavrier P, Knecht DA, Heath C, Stahl P, Grinstein S.

2005 Phosphatidylinositol-4,5-bisphosphate hydrolysis directs actin remodeling during phagocytosis.

J. Cell Biol.

11 , 139 – 149. (doi:10.

1083/jcb.200412162)

60. Welch MD, Rosenblatt J, Skoble J, Portnoy DA,

Mitchison TJ. 1998 Interaction of human Arp2/3 complex and Listeria monocytogenes ActA protein in actin filament elongation.

Science 281 , 105 – 108.

(doi:10.1126/science.281.5373.105)

61. Sagot I, Rodal AA, Moseley J, Goode BL, Pellman D.

2002 An actin nucleation mechanism mediated by

Bni1 and profilin.

Nat. Cell Biol.

4 , 626 – 631.

(doi:10.1038/ncb83)

62. Pruyne D, Evangelista M, Yang C, Bi E, Zigmond S,

Bretscher A, Boone C. 2002 Role of formins in actin assembly: nucleation and barbed-end association.

Science 297 , 612 – 615. (doi:10.1126/science.

1072309)

63. Quinlan ME, Heuser JE, Kerkhoff E, Mullins RD. 2005

Drosophila Spire is an actin nucleation factor.

Nature

433 , 382 – 388. (doi:10.1038/nature03241)

64. Ahuja R, Pinyol R, Reichenbach N, Custer L,

Klingensmith J, Kessels MM, Qualmann B. 2007

Cordon-bleu is an actin nucleation factor and controls neuronal morphology.

Cell 131 , 337 – 350.

(doi:10.1016/j.cell.2007.08.030)

65. Chereau D, Boczkowska M, Skwarek-Maruszewska A,

Fujiwara I, Hayes DB, Rebowski G, Lappalainen P,

Pollard TD, Dominguez R. 2008 Leiomodin is an actin filament nucleator in muscle cells.

Science

320 , 239 – 243. (doi:10.1126/science.1155313)

66. Machesky LM, Atkinson SJ, Ampe C, Vandekerckhove

J, Pollard TD. 1994 Purification of a cortical complex containing two unconventional actins from

Acanthamoeba by affinity chromatography on profilin-agarose.

J. Cell Biol.

127 , 107 – 115. (doi:10.

1083/jcb.127.1.107)

67. Lees-Miller JP, Henry G, Helfman DM. 1992

Identification of act2, an essential gene in the fission yeast Schizosaccharomyces pombe that encodes a protein related to actin.

Proc. Natl Acad.

Sci. USA 89 , 80 – 83. (doi:10.1073/pnas.89.1.80)

68. Schwob E, Martin RP. 1992 New yeast actin-like gene required late in the cell cycle.

Nature 355 ,

179 – 182. (doi:10.1038/355179a0)

69. Fyrberg C, Fyrberg E. 1993 A Drosophila homologue of the Schizosaccharomyces pombe act2 gene.

Biochem. Genet.

31 , 329 – 341. (doi:10.1007/

BF00553175)

70. Welch MD, DePace AH, Verma S, Iwamatsu A,

Mitchison TJ. 1997 The human Arp2/3 complex is composed of evolutionarily conserved subunits and is localized to cellular regions of dynamic actin filaments assembly.

J. Cell. Biol.

138 , 375 – 384.

(doi:10.1083/jcb.138.2.375)

71. Waterston R et al . 1992 A survey of expressed genes in Caenorhabditis elegans .

Nat. Genet.

1 , 114 – 123.

(doi:10.1038/ng0592-114)

72. Robinson RC, Turbedsky K, Kaiser DA, Marchand JB,

Higgs HN, Choe S, Pollard TD. 2001 Crystal structure of Arp2/3 complex.

Science 294 , 1679 – 1684.

(doi:10.1126/science.1066333)

73. Mullins RD, Heuser JE, Pollard TD. 1998 interaction of Arp2/3 complex with actin: nucleation, high affinity pointed end capping, and formation of branching networks of actin filaments.

Proc. Natl

Acad. Sci. USA 95 , 6181 – 6186. (doi:10.1073/pnas.

95.11.6181)

74. Amann K, Pollard TD. 2001 Direct real-time observation of actin filament branching mediated by Arp2/3 complex using total internal reflection fluorescence microscopy.

Proc. Natl Acad. Sci. USA

98 , 15 009 – 15 013. (doi:10.1073/pnas.211556398)

75. Rouiller I et al . 2008 The structural basis of actin filament branching by the Arp2/3 complex.

J. Cell

Biol.

180 , 887 – 895. (doi:10.1083/jcb.200709092)

76. Beltzner CC, Pollard TD. 2004 Identification of functionally important residues of Arp2/3 complex by analysis of homology models from diverse species.

J. Mol. Biol.

226 , 551 – 565. (doi:10.1016/j.

jmb.2003.12.017)

77. Machesky LM, Mullins RD, Higgs HN, Kaiser DA,

Blanchoin L, May RC, Hall ME, Pollard TD. 1999

Scar, a WASp-related protein, activates nucleation of actin filaments by the Arp2/3 complex.

Curr. Biol.

96 , 3739 – 3744. (doi:10.1073/pnas.96.7.3739)

78. Rohatgi R, Ma L, Miki H, Lopez M, Kirchhausen T,

Takenawa T, Kirschner MW. 1999 The interaction between N-WASP and the Arp2/3 complex links

Cdc42-dependent signals to actin assembly.

Cell 97 ,

221 – 231. (doi:10.1016/S0092-8674(00)80732-1)

79. Yarar D, To W, Abo A, Welch MD. 1999 The

Wiskott – Aldrich syndrome protein directs actinbased motility by stimulating actin nucleation with the Arp2/3 complex.

Curr. Biol.

9 , 555 – 558.

(doi:10.1016/S0960-9822(99)80243-7)

80. Derivery E, Gautreau A. 2010 Generation of branched actin networks: assembly and regulation

12

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016 of the N-WASP and WAVE molecular machines.

BioEssays 32 , 119 – 131. (doi:10.1002/bies.

200900123)

81. Goley ED, Rodenbusch SE, Martin AC, Welch MD.

2004 Critical conformational changes in the Arp2/3 complex are induced by nucleotide and nucleation promoting factor.

Mol. Cell.

22 , 269 – 279. (doi:10.

1016/j.molcel.2004.09.018)

82. Zencheck WD, Xiao H, Nolen BJ, Angeletti RH,

Pollard TD, Almo SC. 2009 Nucleotide- and activator-dependent structural and dynamic changes of arp2/3 complex monitored by hydrogen/ deuterium exchange and mass spectrometry.

J. Mol.

Biol.

390 , 414 – 427. (doi:10.1016/j.jmb.

2009.03.028)

83. Rodal AA, Sokolova O, Robins DB, Daugherty KM,

Hippenmeyer S, Riezman H, Grigorieff N, Goode BL.

2005 Conformational changes in the Arp2/3 complex leading to actin nucleation.

Nat. Struct.

Mol. Biol.

12 , 26 – 31. (doi:10.1038/nsmb870)

84. Kreishman-Deitrick M et al . 2005 NMR analyses of the activation of the Arp2/3 complex by neuronal

Wiskott – Aldrich syndrome protein.

Biochemistry

44 , 15 247 – 15 256. (doi:10.1021/bi051065n)

85. Kelly AE, Kranitz H, Dotsch V, Mullins RD. 2006

Actin binding to the central domain of WASP/Scar proteins plays a critical role in the activation of the

Arp2/3 complex.

J. Biol. Chem.

281 , 10 589 –

10 597. (doi:10.1074/jbc.M507470200)

86. Chereau D, Kerff F, Graceffa P, Grabarek Z,

Langsetmo K, Dominguez R. 2005 Actin-bound structures of Wiskott – Aldrich syndrome protein

(WASP)-homology domain 2 and the implications for filament assembly.

Proc. Natl Acad. Sci. USA

102 , 16 644 – 16 649. (doi:10.1073/pnas.

0507021102)

87. Miki H, Sasaki T, Taka Y, Takenawa T. 1998 Induction of filopodium formation by a WASP-related actindepolymerrizing protein N-WASP.

Nature 391 ,

93 – 96. (doi:10.1038/34208)

88. Anton IM, Jones GE, Wandosell F, Geha R, Ramesh

N. 2007 WASP-interacting protein (WIP): working in polymerisation and much more.

Trends. Cell. Biol.

11 , 555 – 562. (doi:10.1016/j.tcb.2007.08.005)

89. Rohatgi R, Ho HY, Kirschner MW. 2000 Mechanism of N-WASP activation by CDC42 and phosphatidylinositol 4, 5-bisphosphate.

J. Cell Biol.

150 , 1299 – 1310. (doi:10.1083/jcb.150.6.1299)

90. Prehoda KE, Scott JA, Mullins RD, Lim WA. 2000

Integration of multiple signals through cooperative regulation of the N-WASP-Arp2/3 complex.

Science

290 , 801 – 806. (doi:10.1126/science.290.5492.80)

91. Higgs HN, Pollard TD. 2000 Activation by Cdc42 and

PIP(2) of Wiskott – Aldrich syndrome protein (WASp) stimulates actin nucleation by Arp2/3 complex.

J. Cell Biol.

150 , 1311 – 1320. (doi:10.1083/ jcb.150.6.1311)

92. Carlier MF et al . 2000 GRB2 links signaling to actin assembly by enhancing interaction of neural

Wiskott – Aldrich syndrome protein (N-WASp) with actin-related protein (ARP2/3) complex.

J. Biol.

Chem.

275 , 21 946 – 21 952. (doi:10.1074/jbc.

M000687200)

93. Rohatgi R, Nollau P, Ho HY. 2001 Nck and phosphatidylinositol4,5-bisphosphate synergistically activate actin polymerization through the N-WASP-

Arp2/3 pathway.

J. Biol. Chem.

276 , 26 448 –

26 452. (doi:10.1074/jbc.M103856200)

94. Ho HY, Rohatgi R, Lebensohn AM, Le M, Li J, Gygi

SP, Kirschner MW. 2004 Toca-1 mediates Cdc42dependent actin nucleation by activating the N-

WASP-WIP complex.

Cell 118 , 203 – 216. (doi:10.

1016/j.cell.2004.06.027)

95. Cory GO, Cramer R, Blanchoin L, Ridley AJ. 2003

Phosphorylation of the WASP-VCA domain increases its affinity for the Arp2/3 complex and enhances actin polymerization by WASP.

Mol. Cell 11 , 1229 –

1239. (doi:10.1016/S1097-2765(03)00172-2)

96. Moseley JB, Sagot I, Manning AL, Xu X, Eck MJ,

Pellman D, Goode BL. 2004 A conserved mechanism for Bni1- and mDia1-induced actin assembly and dual regulation of Bni1 by Bud6 and profilin.

Mol. Biol. Cell 15 , 896 – 907. (doi:10.1091/mbc.

E03-08-0621)

97. Xu Y, Moseley JB, Sagot I, Poy F, Pellman D, Goode

BL, Eck MJ. 2004 Crystal structures of a formin homology-2 domain reveal a tethered dimer architecture.

Cell 116 , 711 – 723. (doi:10.1016/

S0092-8674(04)00210-7)

98. Otomo T, Tomchick DR, Otomo C, Panchal SC,

Machius M, Rosen MK. 2005 Structural basis of actin filament nucleation and processive capping by a formin homology 2 domain.

Nature 433 , 488 – 494.

(doi:10.1038/nature03251)

99. Pring M, Evangelista M, Boone C, Yang C, Zigmond

SH. 2003 Mechanism of formin-induced nucleation of actin filaments.

Biochemistry 42 , 486 – 496.

(doi:10.1021/bi026520j)

100. Zigmond SH, Evangelista M, Boone C, Yang C, Dar

AC, Sicheri F, Forkey J, Pring M. 2003 Formin leaky cap allows elongation in the presence of tight capping proteins.

Curr. Biol.

13 , 1820 – 1823.

(doi:10.1016/j.cub.2003.09.057)

101. Kovar DR, Pollard TD. 2004 Insertional assembly of actin filament barbed ends in association with formins produces piconewton forces.

Proc. Natl

Acad. Sci. USA 101 , 14 725 – 14 730. (doi:10.1073/ pnas.0405902101)

102. Romero S, Le Clainche C, Didry D, Egile C, Pantaloni

D, Carlier MF. 2004 Formin is a processive motor that requires profilin to accelerate actin assembly and associated ATP hydrolysis.

Cell 119 , 419 – 429.

(doi:10.1016/j.cell.2004.09.039)

103. Kovar DR, Harris ES, Mahaffy R, Higgs HN, Pollard

TD. 2006 Control of the assembly of ATP- and ADPactin by formins and profilin.

Cell 124 , 423 – 435.

(doi:10.1016/j.cell.2005.11.038,)

104. Higgs HN, Peterson KJ. 2005 Phylogenetic analysis of the formin homology 2 domain.

Mol. Biol. Cell.

16 , 1 – 13. (doi:10.1091/mbc.E04-07-0565)

105. Li F, Higgs HN. 2003 The mouse Formin mDia1 is a potent actin nucleation factor regulated by autoinhibition.

Curr. Biol.

13 , 1335 – 1340. (doi:10.

1016/S0960-9822(03)00540-2)

106. Li F, Higgs HN. 2005 Dissecting requirements for auto-inhibition of actin nucleation by the formin, mDia1.

J. Biol. Chem.

280 , 6986 – 6992. (doi:10.

1074/jbc.M411605200)

107. Peng J, Wallar BJ, Flanders A, Swiatek PJ, Alberts

As. 2003 Disruption of the Diaphanous-related formin Drf1 gene encoding mDia1 reveals a role for Drf3 as an effector for Cdc42.

Curr. Biol.

13 ,

534 – 545. (doi:10.1016/S0960-9822(03)00170-2)

108. Lammers M, Meyer S, Kuhlmann D, Wittenghofer A.

2008 Specificity of interactions between mDia isoforms and Rho proteins.

J. Biol. Chem.

283 ,

35 236 – 35 246. (doi:10.1074/jbc.M805634200)

109. Lammers M, Rose R, Scrima A, Wittenghofer A.

2005 The regulation of mDia1 by autoinhibition and its release by Rho-GTP.

EMBO J.

24 , 4176 – 4187.

(doi:10.1038/sj.emboj.7600879)

110. Posfay-Barbe KM, Wald ER. 2009 Listeriosis.

Semin.

Fetal Neonatal. Med.

14 , 228 – 233. (doi:10.1016/j.

siny.2009.01.006)

111. Brundage RA, Smith GA, Camilli A, Theriot JA,

Portnoy DA. 1993 Expression and phosphorylation of the Listeria monocytogenes ActA protein in mammalian cells.

Proc. Natl Acad. Sci. USA 90 ,

11 890 – 11 894. (doi:10.1073/pnas.90.24.11890)

112. Leung N, Gianfelice A, Gray-Owen SD, Ireton K. 2013

Impact of the Listeria monocytogenes protein InlC on infection in mice.

Infect. Immun.

81 , 1334 – 1340.

(doi:10.1128/IAI.01377-12)

113. Bakardjiev AL, Stacy BA, Fisher SJ, Portnoy DA. 2004

Listeriosis in the pregnant guinea pig: a model for vertical transmission.

Infect. Immun.

72 , 489 – 497.

(doi:10.1128/IAI.72.1.489-497.2004)

114. Racz P, Tenner K, Me´ro¨ E. 1972 Experimental Listeria enteritis . I. An electron microscopic study of the epithelial phase in experimental Listeria infection.

Lab Invest.

26 , 694 – 700.

115. Marco AJ, Altimira J, Prats N, Lopez S, Dominguez L,

Domingo M, Briones V. 1997 Penetration of Listeria monocytogenes in mice infected by the oral route.

Microb. Pathog.

23 , 255 – 263. (doi:10.1006/mpat.

1997.0144)

116. Bakardjiev AI, Stacy BA, Portnoy DA. 2005 Growth of Listeria monocytogenes in the guinea pig placenta and role of cell-to-cell spread in fetal infection.

J. Infect. Dis.

191 , 1889 – 1897. (doi:10.

1086/430090)

117. LeMonnier A, Autret N, Join-Lambert OF, Jaubert F,

Charbit A, Berche P, Kayal S. 2007 ActA is required for crossing of the fetoplacental barrier by Listeria monocytogenes .

Infect. Immun.

75 , 950 – 957.

(doi:10.1128/IAI.01570-06)

118. Theriot JA, Mitchison TJ, Tilney LG, Portnoy DA.

1992 The rate of actin-based motility of intracellular

Listeria monocytogenes equals the rate of actin polymerization.

Nature 357 , 257 – 260. (doi:10.

1038/357257a0)

119. Zalevsky J, Grigorova I, Mullins RD. 2001 Activation of the Arp2/3 complex by the Listeria ActA protein.

ActA binds two actin monomers and three subunits of the Arp2/3 complex.

J. Biol. Chem.

276 ,

3468 – 3475. (doi:10.1074/jbc.M006407200)

120. Cicchetti G, Maurer P, Wagener P, Kocks C. 1999

Actin and phosphoinositide binding by the ActA protein of the bacterial pathogen Listeria

13

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016 monocytogenes .

J. Biol. Chem.

274 , 33 616 – 33 626.

(doi:10.1074/jbc.274.47.33616)

121. Lauer P, Theriot JA, Skoble J, Welch MD, Portnoy

DA. 2001 Systematic mutational analysis of the amino-terminal domain of the Listeria monocytogenes ActA protein reveals novel functions in actin-based motility.

Mol. Microbiol.

42 , 1163 –

1177. (doi:10.1046/j.1365-2958.2001.02677.x)

122. Chakraborty T et al . 1995 A focal adhesion factor directly linking intracellularly motile Listeria monocytogenes and Listeria ivanovii to the actinbased cytoskeleton of mammalian cells.

EMBO J.

14 , 1314 – 1321.

123. Gertler FB, Niebuhr K, Reinhard M, Wehland J,

Soriano P. 1996 Mena, a relative of VASP and

Drosophila enabled, is implicated in the control of microfilament dynamics.

Cell 87 , 227 – 239. (doi:10.

1016/S0092-8674(00)81341-0)

124. Niebuhr K et al . 1997 A novel proline-rich motif present in ActA of Listeria monocytogenes and cytoskeletal proteins is the ligand for the EVH1 domain, a protein module present in the Ena/VASP family.

EMBO J.

16 , 5433 – 5444. (doi:10.1093/ emboj/16.17.5433)

125. Pistor S, Chakraborty T, Walter U, Wehland J. 1995

The bacterial actin nucleator protein ActA of Listeria monocytogenes contains multiple binding sites for host microfilament proteins.

Curr. Biol.

5 , 517 – 525.

(doi:10.1016/S0960-9822(95)00104-7)

126. Smith GA, Theriot J, Portnoy D. 1996 The tandem repeat domain in the Listeria monocytogenes ActA protein controls the rate of actin-based motility, the percentage of moving bacteria, and the localization of vasodilator-stimulated phosphoprotein and profilin.

J. Cell Biol.

135 , 647 – 660. (doi:10.1083/ jcb.135.3.647)

127. Reinhard M, Giel K, Abel K, Haffner C, Jarchau T,

Hoppe V, Jockusch BM, Walter U. 1995 The prolinerich focal adhesion and microfilament protein VASP is a ligand for profilins.

EMBO J.

14 , 1583 – 1589.

128. Machesky LM, Install RH. 1998 Scar1 and the related Wiskott – Aldrich syndrome protein, WASP, regulate the actin cytoskeleton through the Arp2/3 complex.

Curr. Biol.

8 , 1347 – 1356. (doi:10.1016/

S0960-9822(98)00015-3)

129. Winter D, Lechler T, Li R. 1999 Activation of the yeast Arp2/3 complex by Bee1p, a WASP-family protein.

Curr. Biol.

9 , 501 – 505. (doi:10.1016/

S0960-9822(99)80218-8)

130. Chong R, Swiss R, Briones G, Stone KL, Gulcicek EE,

Agaisse H. 2009 Regulatory mimicry in Listeria monocytogenes actin-based motility.

Cell Host

Microbe 6 , 268 – 278. (doi:10.1016/j.chom.

2009.08.006)

131. Stebbins CE, Galan JE. 2001 Structural mimicry in bacterial virulence.

Nature 412 , 701 – 705. (doi:10.

1038/35089000)

132. Ogawa H, Nakamura A, Nakaya R. 1968

Cinemicrographic study of tissue cell cultures infected with Shigella flexneri .

Jpn J. Med. Sci. Biol.

21 , 259 – 273.

133. Sansonetti PJ, Arondel J, Fontaine A, d’Hauteville H,

Bernardini ML. 1991 OmpB (osmo-regulation) and icsA (cell-to-cell spread) mutants of Shigella flexneri: vaccine candidates and probes to study the pathogenesis of shigellosis.

Vaccine 9 , 416 – 422.

(doi:10.1016/0264-410X(91)90128-S)

134. Egile C, Loisel TP, Laurent V, Li R, Pantaloni D,

Sansonetti PJ, Carlier MF. 1999 Activation of the

CDC42 effector N-WASP by the Shigella flexneri IcsA protein promotes actin nucleation by Arp2/3 complex and bacterial actin-based motility.

J. Cell

Biol.

146 , 1319 – 1332. (doi:10.1083/jcb.146.6.1319)

135. Lommel S, Benesch S, Rottner K, Franz K, Wehland

J, Kuhn R. 2001 Actin pedestal formation by enteropathogenic Escherichia coli and intracellular motility of Shigella flexneri are abolished in

N-WASP-defective cells.

EMBO Rep.

2 , 850 – 857.

(doi:10.1093/embo-reports/kve197)

136. Snapper SB et al . 2001 N-WASP deficiency reveals distinct pathways for cell surface projections and microbial actin-based motility.

Nat. Cell Biol.

3 ,

897 – 904. (doi:10.1038/ncb1001-897)

137. Mounier J, Laurent V, Hall A, Fort P, Carlier MF,

Sansonetti PJ, Egile C. 1999 Rho family GTPases control entry of Shigella flexneri into epithelial cells but not intracellular motility.

J. Cell Sci.

112 ,

2069 – 2080.

138. Shibata T, Takeshima F, Chen F, Alt FW, Snapper SB.

2002 Cdc42 facilitates invasion but not the actinbased motility of Shigella .

Curr. Biol.

12 , 341 – 345.

(doi:10.1016/S0960-9822(02)00689-9)

139. Moreau V, Frischknect F, Reckmann I, Vincentelli R,

Rabut G, Stewart D, Way M. 2000 A complex of

N-WASP and WIP integrates signalling cascades that lead to actin polymerization.

Nat. Cell Biol.

2 ,

441 – 448. (doi:10.1038/35017080)

140. Leung Y, Ally S, Goldberg MB. 2008 Bacterial actin assembly requires toca-1 to relieve N-Wasp autoinhibition.

Cell Host Microbe 3 , 39 – 47. (doi:10.

1016/j.chom.2007.10.011)

141. Gouin E, Gantelet H, Egile C, Lasa I, Ohayon H,

Villiers V, Gounon P, Sansonetti PJ, Cossart P. 1999

A comparative study of the actin-based motilities of the pathogenic bacteria Listeria monocytogenes ,

Shigella flexneri and Rickettsia conorii .

J. Cell. Sci.

112 , 1697 – 1708.

142. Harlander RS, Way M, Ren Q, Howe D, Grieshaber

SS, Heinzen RA. 2003 Effects of ectopically expressed neuronal Wiskott – Aldrich syndrome protein domains on Rickettsia rickettsii actin-based motility.

Infect. Immun.

71 , 1551 – 1556. (doi:10.

1128/IAI.71.3.1551-1556.2003)

143. Serio AW, Jeng RL, Haglund CM, Reed LJ, Welch

MD. 2010 Defining a core set of actin cytoskeletal proteins critical for actin-based motility.

Cell Host

Microbe 7 , 388 – 398. (doi:10.1016/j.chom.

2010.04.008)

144. Engelbrecht F, Chun S-K, Ochs C, Hess J, Lottspeich

F, Goebel W, Sokolovic Z. 1996 A new PrfAregulated gene of Listeria monocytogenes encoding a small, secred protein which belongs to the family of internalins.

Mol. Microbiol.

21 , 823 – 837. (doi:10.

1046/j.1365-2958.1996.541414.x)

145. Otani T, Ichii T, Aono S, Takeichi M. 2006 Cdc42 GEF

Tuba regulates the junctional configuration of simple epithelial cells.

J. Cell. Biol.

175 , 135 – 146.

(doi:10.1083/jcb.200605012)

146. Ivanov AI, Nusrat A, Parkos CA. 2005 Endocytosis of the apical junctional complex: mechanisms and possible roles in regulation of epithelial barriers.

Bioessays 27 , 356 – 365. (doi:10.1002/bies.20203)

147. Miyoshi J, Takai Y. 2005 Molecular perspective on tight-junction assembly and epithelial polarity.

Adv.

Drug. Deliv. Rev.

57 , 815 – 855. (doi:10.1016/j.addr.

2005.01.008)

148. Fehon RG, McClatchey AI, Bretscher A. 2010

Organizing the cell cortex: the role of ERM proteins.

Nat. Rev. Mol. Cell Biol.

11 , 276 – 287. (doi:10.1038/ nrm2866)

149. Turenen O, Wahlstrom T, Vaheri A. 1994 Ezrin had a

COOH-terminal actin binding site that is conserved in the ezrin protein family.

J. Cell Biol.

126 ,

1445 – 1453. (doi:10.1083/jcb.126.6.1445)

150. Gary R, Bretscher A. 1995 Ezrin self-association involves binding of an N-terminal domain to a normally masked C-terminal domain that includes the

F-actin binding site.

Mol. Biol. Cell.

6 , 1061–1075.

151. Reczek D, Berryman M, Bretscher A. 1997

Identification of EBP50: A PDZ-containing phosphoprotein that associates with members of the ezrin-radixin-moesin family.

J. Cell Biol.

139 ,

169 – 179. (doi:10.1083/jcb.139.1.169)

152. Yonemura S, Hirao M, Doi Y, Takahashi N, Kondo T,

Tsukita S, Tsukita S. 1998 Ezrin/radixin/moesin

(ERM) proteins bind to a positively charged amino acid cluster in the juxta-membrane cytoplasmic domain of CD44, CD43, and ICAM-2.

J. Cell Biol.

140 , 885 – 895. (doi:10.1083/jcb.140.4.885)

153. Nakamura F, Amieva MR, Furhmayr H. 1995

Phosphorylation of threonine 558 in the carboxylterminal actin-binding domain of moesin by thrombin activation of human platelets.

J. Biol.

Chem.

270 , 31 377 – 31 385. (doi:10.1074/jbc.270.

20.12319)

154. Yonemura S, Matsui T, Tsukita S. 2002 Rhodependent and -independent activation mechanisms of ezrin/radixin/moesin proteins: an essential role for polyphosphoinositides in vivo .

J. Cell Sci.

115 , 2569 – 2580.

155. Fievet BT, Gautreau A, Roy C, Del Maestro L,

Mangeat P, Louvard D, Arpin M. 2004

Phosphoinositide binding and phosphorylation act sequentially in the activation mechanism of ezrin.

J. Cell Biol.

164 , 653 – 659. (doi:10.1083/jcb.

200307032)

156. Sechi AS, Wehland J, Small JV. 1997 The isolated comet tail pseudopodium of Listeria monocytogenes : a tail of two actin filament populations, long and axial and short and random.

J. Cell. Biol.

137 ,

155 – 167. (doi:10.1083/jcb.137.1.155)

157. Romero S, Van Nhieu GTV. 2009 A bacterial virulence factor that dissipates tension.

Nat. Cell

Biol.

11 , 1174 – 1175. (doi:10.1038/ncb1009-1174)

158. Heindl JE, Saran I, Yi Cr, Lesser CF, Goldberg MB.

2010 Requirement for formin-induced actin polymerization during spread of Shigella flexneri .

Infect. Immun.

78 , 193 – 203. (doi:10.1128/IAI.

00252-09)

14

Downloaded from http://rsob.royalsocietypublishing.org/ on October 2, 2016

159. Cheney RE, Sousa AD. 2005 Myosin-10: a molecular motor at the cell’s fingertips.

Trends Cell Biol.

15 ,

533 – 539. (doi:10.1016/j.tcb.2005.08.006)

160. Sansonetti PJ, Mournier J, Prevost MC, Mege RM.

1994 Cadherin expression is required for the spread of Shigella flexneri between epithelial cells.

Cell 76 ,

829 – 839. (doi:10.1016/0092-8674(94)90358-1)

161. Robbins JR, Barth AI, Marquis H, de Hostos EI,

Nelson WJ, Theriot JA. 1999 Listeria monocytogenes exploits normal host cell processes to spread from cell to cell.

J. Cell. Biol.

146 , 1333 – 1350. (doi:10.

1083/jcb.146.6.1333)

162. Tran Van Nhieu G, Clair C, Bruzzone R, Mesnil M,

Sansonetti PJ, Combettes L. 2003 Connexindependent inter-cellular communication increases invasion and dissemination of Shigella in epithelial cells.

Nat. Cell Biol.

5 , 720 – 726. (doi:10.1038/ ncb1021)

163. Mariano C, Sasaki H, Brites D, Brito MA. 2011 A look at tricellulin and its role in tight junction formation and maintenance.

Eur. J. Immunol.

90 , 787 – 796.

(doi:10.1016/j.ejcb.2011.06.005)

164. Ikenouchi J, Furuse M, Furuse K, Sasaki H, Tsukita S,

Tsukita S. 2005 Tricellulin constitutes a novel barrier at tricellular contacts of epithelial cells.

J. Cell Biol.

171 , 939 – 945. (doi:10.1083/jcb.200510043)

165. Loisel TP, Boujemaa R, Pantaloni D, Carlier M-F.

1999 Reconstitution of actin-based motility of

Listeria and Shigella using pure proteins.

Nature

401 , 613 – 616. (doi:10.1083/jcb.144.6.1245)

166. Gomez GA, McLachlan RW, Yap AS. 2011 Productive tension: force-sensing and homeostasis of cell – cell junctions.

Trends Cell Biol.

21 , 499 – 505. (doi:10.

1016/j.tcb.2011.05.006)

167. Luna A, Matas OB, Martinez-Menarguez JA, Mato E,

Duran JM, Ballesta J, Way M, Egea G. 2002

Regulation of protein transport from the Golgi complex to the endoplasmic reticulum by CDC42 and N-WASP.

Mol. Biol. Cell.

13 , 866 – 879. (doi:10.

1091/mbc.01-12-0579)

168. Matas OB, Martinez-Menarguez JA, Egea G. 2004

Association of Cdc42/N-WASP/Arp2/3 signaling pathway with Golgi membranes.

Traffic 5 , 838 –

846. (doi:10.1111/j.1600-0854.2004.00225.x)

169. Harris K, Tepass U. 2010 Cdc42 and vesicle trafficking in polarized cells.

Traffic 11 , 1272 – 1279.

(doi:10.1111/j.1600-0854.2010.01102.x)

170. Salazar MA et al . 2003 Tuba, a novel protein containing Bin/Amphiphysin/Rvs and Dbl homology domains, links dynamin to regulation of the actin cytoskeleton.

J. Biol. Chem.

278 , 49 031 – 49 043.

(doi:10.1074/jbc.M308104200)

171. Kodani A, Kristensen I, Huang L, Sutterlin C. 2009

GM130-dependent control of Cdc42 activity at the

Golgi regulates centrosome organization.

Mol. Biol.

Cell.

20 , 1192 – 1200. (doi:10.1091/mbc.

E08-08-0834)

172. Mim C, Unger VM. 2012 Membrane curvature and its generation by BAR proteins.

Trends Biochem. Sci.

37 , 526 – 533. (doi:10.1016/j.tibs.2012.09.001)

173. Yap AS, Crampton MS, Hardin J. 2007 Making and breaking contacts: the cellular biology of cadherin regulation.

Curr. Opin. Cell Biol.

19 , 508 – 514.

(doi:10.1016/j.ceb.2007.09.008)

174. Wirtz-Peitz F, Zallen JA. 2009 Junctional trafficking and epithelial morphogenesis.

Curr. Opin. Genet.

Dev.

19 , 350 – 356. (doi:10.1016/j.gde.2009.04.011)

175. de Beco S, Gueudry C, Amblard F, Coscoy S. 2009

Endocytosis is required for E-cadherin redistribution at mature adherens junctions.

Proc. Natl Acad. Sci.

USA 106 , 7010 – 7015. (doi:10.1073/pnas.

0811253106)

176. Nejsum LN, Nelson WJ. 2009 Epithelial cell surface polarity: the early steps.

Front. Biosci.

14 ,

1088 – 1098. (doi:10.2741/3295)

177. Guo W, He B. 2009 The exocyst complex in polarized exocytosis.

Curr. Opin. Cell Biol.

21 ,

537 – 542. (doi:10.1016/j.ceb.2009.04.007)

178. Schmid SL, Frolov VA. 2011 Dynamin: functional design of a membrane fission catalyst.

Annu. Rev.

Cell. Dev. Biol.

27 , 79 – 105. (doi:10.1146/annurevcellbio-100109-104016)

179. McCullough J, Colf LA, Sundquist WI. 2013

Membrane fission reactions of the mammalian

ESCRT pathway.

Annu. Rev. Biochem.

82 , 663 – 692.

(doi:10.1146/annurev-biochem-072909-101058)

180. Doyle ET, Burns SM, Contag CH. 2004 In vivo bioluminescence imaging for integrated studies of infection.

Cell. Microbiol.

6 , 303 – 317. (doi:10.1111/ j.1462-5822.2004.00378.x)

15

Download